134
Copyright by Chris Eugene Brough 2015

Copyright by Chris Eugene Brough 2015

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Copyright by Chris Eugene Brough 2015

Copyright

by

Chris Eugene Brough

2015

Page 2: Copyright by Chris Eugene Brough 2015

The Dissertation Committee for Chris Eugene Brough Certifies that this is the

approved version of the following dissertation:

Novel Uses of Pharmaceutical Polymers as enabled by KinetiSol®

Dispersing

Committee:

Robert O. Williams III, Supervisor

James W. McGinity

Christopher R. Frei

Sean M. Kerwin

Dave A. Miller

Page 3: Copyright by Chris Eugene Brough 2015

Novel Uses of Pharmaceutical Polymers as enabled by KinetiSol®

Dispersing

by

Chris Eugene Brough, B.S., M.S., M.B.A.

Dissertation

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin

August, 2015

Page 4: Copyright by Chris Eugene Brough 2015

Dedication

To my wife, Jessica, who has been a huge support in all of my business and scholastic

adventures.

To my three kids, Ezra, Scarlett & Levi, who have been patient with their dad as he has

had some late nights and early mornings.

To my father, who always wanted one of his boys to become a ‘doctor’.

Page 5: Copyright by Chris Eugene Brough 2015

v

Acknowledgements

I would like to thank my boss, Gershon Yaniv, for pushing me to pursue this Ph.D.

and for also allowing me to be employed full-time to allow it to happen. I would like to

thank Robert O. Williams III for being such a wonderful supervisor and allowing me the

flexibility to juggle school and work. I could not have completed the program without his

patience, understanding, support and guidance. James W. McGinity’s support of the

technology has not only helped DisperSol get its start, but also sparked interest within me

for this particular field of study, of which I am very grateful.

I would like to thank the faculty and the members on my committee. I cannot take

the time here to write all of the things I learned nor the ways I was influenced, but I am

very grateful for them. This may sound cheesy, but I really felt that every one of them

genuinely cared about my success. In the vigor of this program, it was nice to feel that

everyone was on my side.

I need to specifically thank Dave Miller, Justin Keen and Justin Hughey. As I think

about the hours that these men spent with me, I am humbled by their generosity, knowledge

and kindness. I could not have done this without them. Also, as funny as this sounds, I

want to thank Texas. I know I would not have ever gotten involved in this industry, if my

company wasn’t bought and moved to Texas.

There is no way that I can adequately thank my wife, Jessica. She has had to handle

most parental duties without me. I know that being a ‘single mother’ to our three kids has

been a real challenge at times, but she has been amazing. She is SuperMom and SuperWife,

but the world will never know it. I just want her to know that I do.

Page 6: Copyright by Chris Eugene Brough 2015

vi

Finally, I would like to thank my Father in Heaven. As I was contemplating getting

this Ph.D. (rather as Gershon was pushing me to do it), I really didn’t want to because I felt

that my two master’s degrees were more than sufficient for what I wanted to do in life. As

I made it a matter of prayer, it was clear to me that this was something that God wanted me

to do. I could not count the times that I wanted to quit or the times that I felt bad as I saw

how this program was negatively effecting my wife and kids. But, each time I turned to

Him, I received comfort and strength. I honestly still don’t know why I did this. If I knew

how hard it was going to be, I probably wouldn’t have done it. But, I have faith that there

is a reason. At the end of the day, I just want to be His servant and to make the world a

little better place. I just hope that the sacrifice this bundle of paper represents actually

results in the outcome that He wanted.

Page 7: Copyright by Chris Eugene Brough 2015

vii

Novel Uses of Pharmaceutical Polymers as enabled by KinetiSol®

Dispersing

Chris Eugene Brough, Ph.D.

The University of Texas at Austin, 2015

Supervisor: Robert O. Williams III

Poor water-solubility is a common characteristic of drug candidates in

pharmaceutical development pipelines today. Various processes have been developed to

increase the solubility, dissolution rate and bioavailability of these active ingredients

belonging to BCSII and IV classifications. Over the last decade, nano-crystal delivery

forms and amorphous solid dispersions have become well established in commercially

available products and industry literature. Chapter 1 is a comparative analysis of these two

methodologies primarily for orally delivered medicaments. The thermodynamic and

kinetic theories relative to these technologies are presented along with a survey of

commercial relevant scientific literature. Marketed products from both technologies are

presented, but there appears to be more amorphous dispersion products on the U.S. market

today and current development trends are showing an industry preference for amorphous

solid dispersions.

Many pharmaceutical polymers have been investigated as the primary component

in amorphous solid dispersions for their ability to increase the apparent water solubility of

poorly water-soluble drugs. Polyvinyl alcohol (PVAL) has not been investigated as a

concentration enhancing polymer owing to its high melting point/high viscosity and poor

organic solubility. Due to the unique attributes of the KinetiSol® Dispersing (KSD)

Page 8: Copyright by Chris Eugene Brough 2015

viii

technology, PVAL has been enabled for this application and Chapter 2 contains an initial

investigation into various grades for improvement of the solubility and bioavailability of

the poorly water-soluble model drug, itraconazole (ITZ). Polymer grades were chosen with

variation in molecular weight and degree of hydroxylation to determine the effects on

performance. Differential scanning calorimetry, powder x-ray diffraction, polarized light

microscopy, size exclusion chromatography and dissolution testing were used to

characterize the amorphous dispersions. An in vivo pharmacokinetic study in rats was also

conducted to compare the selected formulation to current market formulations of ITZ.

Chapter 3 continues the investigation into the use of PVAL as a concentration

enhancing polymer for amorphous solid dispersion. The previous chapter revealed that the

88% hydrolyzed grade was optimal for ITZ compositions with regard to solid-state

properties, non-sink dissolution performance and bioavailability enhancement. This

chapter explores the influence of molecular weight for the 88% hydrolyzed grade in the

range of 4 to 8 mPa·s with the top performing grade from both chapters emerging as PVAL

4-88. Amorphous dispersions at 10, 20, 30, 40 and 50% ITZ drug loads in PVAL 4-88

were compared by dissolution performance. Analytical tools of diffusion-ordered

spectroscopy and Fourier transform infrared spectroscopy were employed to understand

the interaction between drug and polymer. Finally, results from a 30 month stability test

of a 30% drug loaded ITZ:PVAL 4-88 composition shows that stable amorphous

dispersions can be achieved.

The KinetiSol® Dispersing (KSD) technology has been shown to create solid

dispersion systems from challenging drugs and highly viscous polymers. The focus has

been primarily using this technology for solubility enhancement, but it can be advantageous

for other obstacles facing the pharmaceutical development industry. Chapter 4 contains an

investigation into the use of the technology for producing abuse deterrent formulations for

Page 9: Copyright by Chris Eugene Brough 2015

ix

the drug, theophylline, which is used as a model for oxycodone. Various high molecular

weight polymers are combined with plasticizers to produce mechanical and chemical

properties sufficient to resist alcohol dose dumping, size reduction for immediate release

and syringeability for injection. Thus, the KinetiSol® Dispersing (KSD) technology can be

used as a formulation platform for creating abuse deterrent delivery forms in addition to

solid amorphous dispersions for solubility enhancement.

Page 10: Copyright by Chris Eugene Brough 2015

x

Table of Contents

List of Tables ...................................................................................................... xiii

List of Figures ..................................................................................................... xiv

List of Equations ................................................................................................ xvi

Chapter 1: Amorphous Solid Dispersions and Nano-crystal Technologies for

Poorly Water-Soluble Drug Delivery ........................................................17

1.1 INTRODUCTION.........................................................................................17

1.2 SCIENTIFIC FOUNDATION .........................................................................18

1.2.1 Nano-crystal Delivery Forms ......................................................19

1.2.2 Amorphous Dispersions ..............................................................22

1.3 PROCESSES ...............................................................................................26

1.3.1 Nano-crystal Technologies .........................................................26

1.3.2 Amorphous Dispersion Technologies .........................................27

1.4 RESEARCH COMPARISONS ........................................................................30

1.5 MARKETED PRODUCTS .............................................................................36

1.5.1 Nano-crystal Products .................................................................36

1.5.2 Amorphous Dispersion Products ................................................38

1.6 CONCLUSION ............................................................................................44

Chapter 2: New Use of Polyvinyl Alcohol as a Solubility Enhancing Polymer for

Poorly Water-soluble Drug Delivery (Part 1) ..........................................45

2.1 Introduction .............................................................................................45

2.2 Materials .................................................................................................51

2.3 Methods...................................................................................................51

2.3.1 KinetiSol® Dispersing (KSD) .....................................................51

2.3.2 Differential Scanning Calorimetry (DSC) ..................................51

2.3.3 Powder X-Ray Diffraction (XRD) ..............................................52

2.3.4 Polarized Light Microscopy (PLM) ............................................52

2.3.5 Size Exclusion Chromatography (SEC)......................................52

2.3.6 Non-sink Dissolution Analysis ...................................................53

Page 11: Copyright by Chris Eugene Brough 2015

xi

2.3.7 High Performance Liquid Chromatography ...............................53

2.3.8 In Vivo Studies ............................................................................53

2.4 Results and Discussion ...........................................................................55

2.4.1 Preliminary Evaluation of PVAL Molecular Weight .................55

2.4.2 Solid-state Analysis ....................................................................57

2.4.3 Non-sink, Gastric Transfer Dissolution Analysis .......................65

2.4.4 In Vivo Study...............................................................................67

2.5 Conclusion ..............................................................................................70

Chapter 3: New Use of Polyvinyl Alcohol as a Solubility Enhancing Polymer for

Poorly Water-soluble Drug Delivery (Part 2) ..........................................71

3.1 Introduction .............................................................................................71

3.2 Materials .................................................................................................74

3.3 Methods...................................................................................................74

3.3.1 KinetiSol® Dispersing (KSD) .....................................................74

3.3.2 Powder X-Ray Diffraction (XRD) ..............................................74

3.3.3 Polarized Light Microscopy (PLM) ............................................75

3.3.4 High Performance Liquid Chromatography (HPLC) .................75

3.3.5 Non-Sink Dissolution Analysis...................................................75

3.3.6 Precipitation Inhibition Analysis ................................................76

3.3.7 Fourier Transform Infrared Spectroscopy (FTIR) ......................76

3.3.8 Diffusion-Ordered Spectroscopy (DOSY) ..................................77

3.4 Results and Discussion ...........................................................................78

3.4.1 Evaluation of PVAL Molecular Weight .....................................78

3.4.2 Drug Loading Evaluation ............................................................80

3.4.3 Evaluation of the ITZ/PVAL Interaction ....................................83

3.4.4 Evaluation of Stability ................................................................88

3.5 Conclusion ..............................................................................................92

Chapter 4: Novel formulation approach for multiparticulate abuse deterrent

development .................................................................................................93

4.1 Introduction .............................................................................................93

Page 12: Copyright by Chris Eugene Brough 2015

xii

4.2 Materials .................................................................................................99

4.3 Methods...................................................................................................99

4.3.1 KinetiSol® Dispersing (KSD) .....................................................99

4.3.2 Non-Sink Dissolution Analysis.................................................100

4.3.3 Dart Impact Test .......................................................................100

4.3.4 Coffee Grinder Size Reduction Test .........................................100

4.3.5 Syringeability Test ....................................................................101

4.3.6 Water Extraction Test ...............................................................101

4.4 Results and Discussion .........................................................................102

4.4.1 Formulation Development ........................................................102

4.4.2 Abuse Testing ...........................................................................105

4.5 Conclusion ............................................................................................113

Bibliography .......................................................................................................114

Vita .....................................................................................................................134

Page 13: Copyright by Chris Eugene Brough 2015

xiii

List of Tables

Table 1.1: Examples of commercial nano crystal products [58] ........................36

Table 1.2: Examples of current amorphous dispersion products [81] (modified).

...........................................................................................................38

Table 2.1: Properties of PVAL (EMD Millipore) [109] ....................................47

Table 2.2: Rat PK study design ..........................................................................54

Table 2.3: Rat PK mean calculations .................................................................69

Table 4.1: Categories for abuse deterrence as defined by the U.S. FDA. ..........94

Table 4.2: Formulations tested to determine excipient trends (n=1). ..............103

Page 14: Copyright by Chris Eugene Brough 2015

xiv

List of Figures

Figure 1.1: Three basic quantities that determine solubility of solid solute. .......22

Figure 1.2: ‘Spring & Parachute’ concept for supersaturation stability [43]

(adapted). ..........................................................................................24

Figure 1.3: Dose proportionality of compound 3 in 10% Tween (-▲-) as the

crystalline form; in 0.5% Methocel/0.24% SLS as the amorphous form (-

■-) and in solid dispersion at 50% drug loading in HPMC-AS given as

suspension in Methocel (-●-) from 10 mpk to 750 mpk dosed at 5 ml/kg

in Sprague Dawley rats (n=4) [74] (corrected). ................................33

Figure 1.4: Timeline of FDA approved amorphous dispersion and nano-crystal

products. ............................................................................................42

Figure 2.1: SEC profiles of unprocessed powders and processed compositions.56

Figure 2.2: XRD profiles of PVAL grades used in study....................................58

Figure 2.3: XRD profiles of ITZ and KSD compositions ...................................58

Figure 2.4: Nonreversing Heat Flow DSC thermograms for 4 mPa·s grades of PVAL

...........................................................................................................60

Figure 2.5: Reversing Heat Flow DSC thermograms for 4 mPa·s grades of PVAL

...........................................................................................................60

Figure 2.6: Nonreversing Heat Flow DSC thermograms for KSD compositions and

ITZ ....................................................................................................61

Figure 2.7: Reversing Heat Flow DSC thermograms for KSD compositions and ITZ

...........................................................................................................62

Figure 2.8: Dissolution Profiles for 4 mPa·s grades KSD compositions.............66

Page 15: Copyright by Chris Eugene Brough 2015

xv

Figure 2.9: Dissolution Profiles for 4 mPa·s grades KSD compositions presented as

concentration (C) relative to the saturation concentration (Cs) of ITZ at

the respective pH...............................................................................66

Figure 2.10: Rat PK of PVAL 4-88 compositions compared with Onmel™. .......68

Figure 3.1: Dissolution profiles of 20% ITZ in PVAL 4-, 5-, 8- and 18-88. ......78

Figure 3.2: Precipitation inhibition curves of 20% ITZ in PVAL 4-88, 5-88, 8-88

and 18-88. .........................................................................................79

Figure 3.3: Dissolution comparison of different ITZ loadings in PVAL 4-88. ...81

Figure 3.4: FTIR-ATR spectra of amorphous ITZ, PVAL 4-88, and amorphous

dispersions.........................................................................................84

Figure 3.5: DOSY maps of (A) PVAL 4-88 and (B) 20% ITZ in PVAL 4-88

amorphous dispersion. ......................................................................86

Figure 3.6: XRD profiles of ITZ, initial and 30 month old amorphous dispersions.

...........................................................................................................89

Figure 3.7: pH change dissolution comparing initial and 30 month ambient storage

amorphous dispersions of ITZ:PVAL 4-88 (30% drug loading). .....90

Figure 4.1: Dissolution profiles of lead formulation in 0.1 N HCl and 40% Ethanol

(n=3). ...............................................................................................105

Figure 4.2: Post Coffee Grinder Test Particle Size distribution. .......................106

Figure 4.3: Dissolution Profile of Original Size and Coffee Ground Material (n=3).

.........................................................................................................107

Figure 4.4: Mortar and Pestle approach for size reduction of multiparticulates.108

Figure 4.5: Result of Syringeability test of lead formulation. ...........................109

Figure 4.6: Dissolution Profiles from the Water Extraction Test. .....................111

Page 16: Copyright by Chris Eugene Brough 2015

xvi

List of Equations

Equation 1.1: Maximum absorbable dose (MAD) formula...............................18

Equation 1.2: The Nernst-Brunner equation. ....................................................19

Equation 1.3: Ostwald-Freundlich equation. .....................................................20

Equation 1.4: Kelvin equation. ..........................................................................21

Equation 1.5: Three basic quantities governing solubility. ...............................22

Page 17: Copyright by Chris Eugene Brough 2015

17

Chapter 1

Amorphous Solid Dispersions and Nano-crystal Technologies for Poorly

Water-Soluble Drug Delivery [1]

[1] Previously published: Brough, C. and R.O. Williams III, Amorphous solid

dispersions and nano-crystal technologies for poorly water-soluble drug delivery.

International Journal of Pharmaceutics, 2013. 453(1): p. 157-166.

Research and writing was supervised by Robert O. Williams III.

1.1 INTRODUCTION

Poorly-water soluble drug substances are a significant percentage of the molecular

entities in the industry’s drug development pipeline and are a growing percentage of those

commercially available [2-4]. In the past, the industry consensus was to view these as

highly risky development candidates [5]. However, given their prevalence, industry

consensus has shifted from an attitude of avoidance to one of acceptance as increasing

research dedication is given to solving solubility challenges [6]. Whether the increase in

number of poorly water-soluble entities is due to modern high-through put screening

methodologies [7] or credited to development on increasingly biocomplex diseases

requiring higher lipophilicity and larger molecular weight, the challenge of poor solubility

is not disappearing in the foreseeable future [8, 9].

Addressing this issue, the pharmaceutical industry has developed multiple methods

for increasing the apparent solubility of crystalline drugs. Traditionally, salt formation was

preferred by medicinal and synthetic chemists for weak bases or weak acids [10].

Unfortunately, only 20-30% of new molecules form salts easily, so 70-80% of those entities

must find another route to improved solubility [11]. Cyclodextrins [12], self-emulsifying

drug delivery systems (SEDDS)[13], solid lipid nanoparticles [14], liposomes [15],

micelles [16], soft gelatin capsules [17], co-crystals [18], pH micro-environmental

Page 18: Copyright by Chris Eugene Brough 2015

18

modifiers [19], and high energy polymorphs [20] are some of the routes available to

pharmaceutical scientists. However, the last decade has shown the prevalence of two

solubility/dissolution rate improvement methods both in scientific literature and marketed

products: nano-crystal delivery forms and solid amorphous dispersions.

1.2 SCIENTIFIC FOUNDATION

Solubility of the drug substances plays a significant role in its bioavailability. The

Biopharmaceutical classification system (BCS) was established in order to classify

intestinal absorption [21]. Poorly water-soluble compounds are grouped into BCS Class II

(high permeability, low solubility) or BCS Class IV (low permeability, low solubility)

depending upon their permeability categorization [22].

There are multiple complex factors in drug absorption, but a straightforward

conceptual approach to understanding solubility’s role can be expressed by the maximum

absorbable dose (MAD) formula:

MAD = 𝐾𝑎 ∙ 𝑆𝑝𝐻 ∙ 𝑉𝑆𝐼 ∙ 𝑡

Equation 1.1: Maximum absorbable dose (MAD) formula.

Where Ka is the intestinal absorption rate constant (related to permeability), SpH is

the solubility at intestinal pH, VSI is the volume of fluid in the small intestine available for

drug dissolution and t is the transit time through the small intestine [23, 24]. While this is

a more simplistic model, it does illustrate important points in the challenge of poorly

soluble drug absorption. First, the solubility at the site of absorption is critical after the pH

transition from the acidic gastric environment [25]. Delivery forms must either include

dissolution in the gastric environment and maintenance of solubility through the pH

transition into the intestines or acceptably rapid dissolution rates at intestinal pH to achieve

Page 19: Copyright by Chris Eugene Brough 2015

19

solubility during the intestinal transit time. Second, Ka, the intestinal absorption rate, must

also be considered with increased solubility. At low solubility, the SpH is likely the limiting

factor in the maximum absorbable dose. If the drug solubility is increased substantially,

the limiting factor can shift from solubility to the intestinal absorption rate [26]. Thus, it

may be more important to stabilize drug solubility for the entire intestinal transit time rather

than to maximize solubility for a portion of it.

1.2.1 Nano-crystal Delivery Forms

Both nano-crystal delivery forms and solid amorphous dispersions increase the

amount of drug dissolved at the site of absorption, but they achieve this by different

mechanisms. Nano-crystal delivery relies on reduced particle size for increased solubility

and dissolution rate [27]. The Nernst-Brunner equation (Noyes-Whitney equation

modified with Fick’s second law) illustrates the effect of smaller particle size on dissolution

as seen in Equation 1.2:

𝑑𝐶

𝑑𝑡=

𝐷𝑆

𝑉ℎ(𝐶𝑠 − 𝐶)

Equation 1.2: The Nernst-Brunner equation.

Where dC/dt is the change of concentration over time (dissolution), D is the

diffusion coefficient, S is the surface area, V is the volume of the dissolution medium, h is

the thickness of the diffusion layer, CS is the saturation solubility and C is the instantaneous

concentration at time t [28]. As particles are reduced in size the surface area increases; as

particles are reduced to the nano scale the surface area increases dramatically. In addition

to affecting surface area, with particle sizes less than about 50µm the thickness of the

diffusion layer appears to decreases as well [29]. Thus, nano sizing increases dissolution

Page 20: Copyright by Chris Eugene Brough 2015

20

by simultaneously increasing surface area in the numerator and decreasing the diffusion

layer thickness in the denominator of Equation 1.2.

The relationship of solubility of a small particle in a bulk solution is given by the

Ostwald-Freundlich equation (Equation 1.3), which is analogous to the Gibbs-Kelvin

equation [30].

ln (𝑥𝑅

𝑥∞) =

2σαγυ2γ

𝑘𝐵𝑅𝑇

Equation 1.3: Ostwald-Freundlich equation.

Where the solubility, xR, of a small solid particle (phase γ) in the ideal bulk solution

(phase α) is related to radius, R. The other variables are as follows: x∞ is the relative

concentration of solute in phase α, σαγ is the surface tension of the solid particle at its

boundary with phase α, υ2γ is the volume per molecule in the solid particle, kB is the

Boltzmann constant and T is the temperature. According to the equation, particles with a

very small radius will have increased solubility. This indicates that nanoparticles not only

affect surface area, S, and the diffusion layer thickness, h, in the Nernst-Brunner equation,

but also the saturation solubility, CS. Entering values to simulate a drug in intestinal fluid

(assuming a drug molecular weight of 500 and an σαγ value of 15-20 mN m-1 for the crystal-

intestinal fluid surface tension) Equation 1.3 predicts an approximate 10-15% increase in

solubility for a particle size of 100 nm [31].

Further illustrations of increased solubility of small particles have been reported

[32, 33]. Junghanns et al. used an equation related to the Ostwald-Freundlich equation, the

Kelvin equation, which describes the relationship between the radius of a liquid droplet

and the vapor pressure leading to evaporation. Equation 1.4 is the Kelvin equation.

Page 21: Copyright by Chris Eugene Brough 2015

21

ln (𝑝

𝑝o) =

2γ𝑉𝑚

Ȓ𝑅𝑇

Equation 1.4: Kelvin equation.

Where p is the vapor pressure, po is the saturated vapor pressure, γ is the surface

tension, Vm is the molar volume, Ȓ is the universal gas constant, R is the radius of the droplet

and T is the temperature. Under the assumption that a transfer of molecules from a liquid

phase to a gas phase is in principal identical to the transfer of molecules from a solid phase

to a liquid phase, the properties of BaSO4 were entered into the Equation 1.4. Using the

curvature of the nanoparticle surfaces to estimate pressures, the Kelvin equation shows an

increase in saturation solubility for particles smaller than 1 µm.

Nanoparticles are much more unstable than microparticles because of the extra

Gibbs free energy contribution related to reducing particle size and primarily due to surface

energy [34]. Addressing this extra contribution is key to formulating pharmaceutical

nanoparticles because they will tend to agglomerate to minimize their total energy [35].

Approaches to stabilizing drug nanoparticles can be categorized into two groups:

thermodynamic stabilization which uses surfactants or block copolymers for particle

stability or kinetic stabilization which uses energy input to compensate for Gibbs free

energy [34]. For maximum effectiveness, the two approaches are often combined [36].

Careful selection of the amount of stabilizer is as important as the selection of the type of

stabilizer. For example, one obstacle to stabilization is Ostwald ripening, which is the

phenomenon in which smaller particles in solution dissolve and deposit on larger particles

in order to reach a more thermodynamically stable state by minimizing the surface to area

Page 22: Copyright by Chris Eugene Brough 2015

22

ratio [37]. Too little stabilizer allows agglomeration of nanoparticles and too much

stabilizer promotes Ostwald ripening [38].

1.2.2 Amorphous Dispersions

Understanding the benefit of solid amorphous dispersions must be explained in

terms of enthalpic energy. There are three basic quantities governing the solubility (S) of

a given solid solute [8]:

𝑆 = 𝑓(𝐶𝑟𝑦𝑠𝑡𝑎𝑙 𝑃𝑎𝑐𝑘𝑖𝑛𝑔 𝐸𝑛𝑒𝑟𝑔𝑦 + 𝐶𝑎𝑣𝑖𝑡𝑎𝑡𝑖𝑜𝑛 𝐸𝑛𝑒𝑟𝑔𝑦 + 𝑆𝑜𝑙𝑣𝑎𝑡𝑖𝑜𝑛 𝐸𝑛𝑒𝑟𝑔𝑦)

Equation 1.5: Three basic quantities governing solubility.

The crystal packing energy term accounts for the energy necessary to disrupt the

crystal lattice and remove isolated molecules. The cavitation energy term accounts for the

energy required to disrupt water in order to create a cavity in which to host the solute

molecule. The solvation energy term accounts for the release of energy as favorable

interactions are formed between the solvent and solute. A graphical representation is

shown in Figure 1.1.

Figure 1.1: Three basic quantities that determine solubility of solid solute.

Page 23: Copyright by Chris Eugene Brough 2015

23

In relative terms, the crystal packing energy is larger than both cavitation and

solvation energies and thus, the driving force behind solubility. The intent in formulating

an amorphous solid dispersion is to minimize this energy component by disrupting the drug

crystal lattice in the delivery form. In conjunction with solubility enhancing polymeric

carriers, the apparent solubility can be increased by up to 10,000 fold or more [39-42]. In

terms of the Nernst-Brunner equation, drastic increases of the saturation solubility, CS,

result in a much faster dissolution rate.

Solid amorphous dispersions essentially contain stored potential energy that can

‘spring’ the molecular entity into a supersaturated state [43, 44]. Since supersaturation is

thermodynamically unstable, the formulation must also provide a ‘parachute’ to keep the

solubility from rapidly returning to the crystalline drug equilibrium solubility (Figure 1.2)

to maintain elevated drug concentrations for the duration of intestinal transit to achieve the

maximum absorbable dose. In general, solubility enhancing polymers are reported to

function well as a parachute or stabilizer due to drug-polymer interactions in solution and

by adsorption of the polymer on the growing crystal [45]. Other solubility enhancing

formulations may require additional excipients to function as the parachute to retard the

descent from high to low energy forms of the drug. Examples of polymers that have been

investigated for their stabilizing effect are PVP [46], PEG [47], methylcellulose (MC) [48]

and HPMC [49].

Page 24: Copyright by Chris Eugene Brough 2015

24

Figure 1.2: ‘Spring & Parachute’ concept for supersaturation stability [44] (adapted).

A concern with solid dispersions is the possibility of the amorphous drug substance

undergoing crystallization on storage. The effect of moisture on storage stability is an

another concern because the presence of water may increase drug mobility and promote

drug crystallization [50]. Additionally, some polymers used in solid dispersions are

hygroscopic, which may result in phase separation, crystal growth or conversion from a

metastable crystalline form to a more stable crystalline structure during storage [51]. This

would result in continually decreasing the solubility and dissolution rate as well as lower

the in vivo performance during the product’s shelf-life.

The above challenges can be mitigated by proper polymer selection, drug loading

optimization and appropriate product packaging selection. Amorphous solid dispersions

can be rendered physically stable via kinetic stabilization: i.e., freezing the amorphous drug

substance in the polymer matrix to restrict molecular mobility to prevent nucleation and

crystal growth. It has been reported that for adequate kinetic stability, the Tg of the

composite matrix should be 50 °C above the maximum storage temperature [52]. From

this perspective, polymer selection is important to ensure a high composite Tg for the

Page 25: Copyright by Chris Eugene Brough 2015

25

preservation of the amorphous drug. A polymer may also stabilize an amorphous drug

substance via drug-polymer intermolecular interactions such as hydrogen bonding, Van der

Waals forces, etc. Such interactions can be estimated a priori utilizing calculated solubility

parameters or empirically via the use of analytical techniques such as Fourier transform

infrared spectroscopy (FTIR) [53-55]. These interactions provide thermodynamic stability

to the amorphous drug substance and can result in product stability irrespective of Tg [56].

Drug loading can impact both kinetic and thermodynamic stability of an amorphous solid

dispersion. For a low Tg drug in a kinetically stabilized dispersion, the higher the drug

loading means the lower the composite Tg, and hence the more unstable the composition.

For a thermodynamically stabilized amorphous solid dispersion, increasing drug loading

can saturate bonding sites, and thus result in a less stable amorphous dispersion. Practically

speaking, most amorphous solid dispersions are formulated in a metastable region where

the mode of stabilization is a combination of both kinetic and thermodynamic mechanisms

[57]. As drug loading impacts both of these stabilizing mechanisms it is critical to conduct

accelerated stability studies on a range of drug loadings to arrive at a physically stable

amorphous dispersion with an acceptable drug load. Finally, packaging is critical when

the amorphous solid dispersion is susceptible to destabilization by moisture absorption

[58]. Adsorbed moisture can lead to crystallization by plasticizing the polymer matrix and

increasing the molecular mobility, displacing drug substances from bonding sites on the

polymer, or both. Utilizing the appropriate dosage form, the packaging configuration can

eliminate moisture contact and stabilize the amorphous product for an acceptable shelf-

life.

Page 26: Copyright by Chris Eugene Brough 2015

26

1.3 PROCESSES

1.3.1 Nano-crystal Technologies

Wet ball milling (also called bead milling or pearl milling) is the most frequently

used production method for drug nano-crystals in the pharmaceutical industry [59]. This

can be attributed in part to the simplicity of the process allowing it to be performed in

almost every lab. The simplest way of doing ball milling is feeding coarse drug substance

into a jar filled with milling media with at least one stabilizing agent. Then, agitate the

milling media by magnetic stirrer or by rotating or tumbling the jar. This will generally

yield very fine particles with a narrow size distribution when allowed to operate long

enough.

Alternatively, wet ball milling operations based on higher energy media movement

can accomplish particle size reduction in times more suitable for industrial pharmaceutical

applications [60]. The NanoCrystal™ process is a high energy wet ball milling process

regarded as the standard procedure to produce nanosuspensions [61]. To operate, the

milling chamber is filled with milling media, water, drug and stabilizer. A shaft with

projectiles spins within the milling chamber which allows drug particles to collide with the

chamber wall, milling media and other drug particles. The high shear forces provide the

energy input to fracture drug crystals into nanometer-size particles. Normal processing

times ranging from 30 minutes to 2 hours yield nanosuspensions of good quality. The

process has demonstrated scalability; batch mode R&D equipment can process 10 mg of

drug substance and larger continuous mode equipment is being used to produce

commercial products [62]. Because it is part of the process, one inherent benefit of wet

ball milling is creating a very stable aqueous suspension. This suspension can be directly

formed into liquid oral, injectable and nebulized inhalation delivery forms [63].

Page 27: Copyright by Chris Eugene Brough 2015

27

High pressure homogenization can be considered the second most frequently used

technique to produce nano-crystals [59]. Like wet ball milling, it is a particle size reduction

technology, but uses jet-stream homogenization by pumping drug, dispersion medium,

surfactants and/or stabilizers under high pressure through a micro fluidizing nozzle. The

particle size reduction is caused by cavitation forces, shear forces and collision through

multiple homogenization cycles. The number of passes depends upon many factors

comprising the type of homogenizer and process conditions, which has led to various

technologies: IDD-P™, Dissocubes® and Nanopure® [32].

Combination technologies have also been developed that integrate a pre-treatment

step with a subsequent high energy step, like high pressure homogenization. The

NANOEDGE™ technology is one example that combines a first classical precipitation

step with a subsequent annealing step by applying high energy (high pressure

homogenization)[32]. The term annealing was used meaning nanoparticles achieve a lower

surface energy by application of energy followed by thermal relaxation. Other combination

technologies, bottom-up methodologies and other processes are available to produce nano-

crystal drug delivery forms [64-66].

1.3.2 Amorphous Dispersion Technologies

Processes for the preparation of solid amorphous dispersions can be categorized

into two general types: solvent methods and fusion or melting methods. With solvent

methods, solid dispersions are obtained by evaporating a common solvent from a drug and

carrier solution. In general, fusion methods heat a drug and carrier composition above their

melting or glass transition temperatures, mix at the elevated temperature, and then cool the

composition in such a way to keep the active ingredient in its amorphous state. Operating

temperatures for solvent techniques are generally lower than fusion techniques and are

Page 28: Copyright by Chris Eugene Brough 2015

28

advantageous for thermolabile drug substances. However, finding a common solvent is

not always straightforward. For example, in the case for solid dispersions for immediate

drug release, hydrophobic drugs are typically combined with hydrophilic carriers, which

can limit solvent selection. A secondary drying step is often required to reduce residual

solvent below accepted levels for safety issues. In addition, small amounts of residual

solvent could negatively affect drug chemical stability and can plasticize the solid

dispersion matrix to subsequently impact physical stability [45].

Practical applications of the solvent method are spray drying and freeze drying. In

spray drying, the drug and carrier solution is atomized into hot gas that causes the solvent

to evaporate resulting in spherical particles containing amorphous drug [67]. Freeze drying

or lyophilization is a technique in which the drug and polymer solution is frozen and the

solvent is sublimed under vacuum. Another solvent method is using a fluidized bed system

to coat multiparticulates (i.e., beads or pellets) with drug-carrier solutions resulting in

pellets with a solid amorphous dispersion coating [68].

Hot-melt extrusion (HME), a well-known fusion technique, has been a topic of

interest in recent pharmaceutical research literature [69]. HME is the process of pumping

compositions through a heated barrel by one or more screws under pressure followed by

discharging the extrudate through a die. Solvents are not necessary thereby eliminating the

aforementioned solvent related issues. The intense mixing and agitation imposed by the

rotating screw(s) causes uniform distribution of drug and the processes is continuous and

efficient. The extrudate discharge is dense and can be post-processed without the bulk

density issues of some solvent methods. Viscous compositions may require the aid of a

plasticizing agent to allow the composition to flow through the heated barrel and out the

die without over torqueing the electric motor. Plasticizers increase molecular mobility

Page 29: Copyright by Chris Eugene Brough 2015

29

upon storage and can allow for crystallization if the composition’s glass transition

temperature is not at least 50°C above the storage condition [52].

Other fusion processes have been explored in creating amorphous dispersions.

Spray congealing is a process where molten compositions are atomized into particles of

spherical shape and then cooled to solidification [70]. KinetiSol® Dispersing is a fusion

based process that has been shown to create amorphous dispersions of high temperature,

thermolabile drug substances without degradation [71]. It does not have the same viscosity

limitations as HME and thus compositions are processed without the need of a plasticizer.

Scalability is a non-issue as the equipment is used in non-pharmaceutical industrial

applications, but no current marketed products are produced by this technology.

Page 30: Copyright by Chris Eugene Brough 2015

30

1.4 RESEARCH COMPARISONS

Numerous papers have been published on amorphous solid dispersions and nano-

crystal formulations. For the purposes of this comparison review, only original research

papers that directly compare the two in a formulation development effort are discussed.

Additionally, literature was limited to those studies that included participation and/or

support from a pharmaceutical company as this was considered to be an indicator of the

study’s commercial relevance.

Fakes et al reported on the enhancement of an HIV-attachment inhibitor, BMS-

488043 [72]. The molecule was classified as a BSC class II compound with low aqueous

solubility of 0.04 mg/ml and a permeability of 178 nm/s in a caco2 cell-line model. To

increase bioavailability of the molecule, a nanosuspension formulation was developed and

compared to an amorphous solid dispersion made by a solvent evaporation method. The

nanosuspension formula was produced by a Nanomill™ at a 10% (w/w) concentration

using hydroxypropyl cellulose (HPC-SL) at 2% (w/w) and sodium lauryl sulfate (SLS) at

0.1% (w/w). The mean cumulative particle size was 0.120 µm and the suspension was

physically stable when stored at room temperature up to 4 weeks.

The amorphous dispersion was produced by flash evaporation from an acetonitrile

solution in a Buchi Rotovap. Polyvinylpyrrolidone (PVP, K-30) was the selected polymer

with drug loadings of 20 and 40%. The dispersions were determined to be XRD amorphous

following storage at 50 °C up to 3 weeks. Further, an extremely high drug load amorphous

dispersion of 80/20 BMS-488043/PVP was found to remain amorphous after storage at 50

°C for 17 weeks. The two amorphous formulations and a single nanosuspension formula

were compared to wet-milled crystalline drug capsule in a crossover beagle dog study. The

two solid dispersions were dosed as 200 mg tablets with the nanosuspension administered

via a standard gavage tube at the equivalent dose. The nanosuspension showed a 4.7 fold

Page 31: Copyright by Chris Eugene Brough 2015

31

increase in Cmax and 4.6 fold increase in AUC over the capsule. The 20% amorphous

dispersion showed an 18.2 and 7.0 fold increase in Cmax and AUC respectively. The 40%

dispersion showed a 15.7 fold increase in Cmax and 8.7 fold increase in AUC. The authors

concluded that the amorphous dispersions were the superior option for increased

bioavailability of BMS-488043.

Zheng et al reported on the commercial development of LCQ789, a BCS Class II

compound, investigated particle size reduction, amorphous dispersions, lipid based

formulations and co-crystals [73]. LCQ789 is a crystalline, neutral compound with a

molecular weight of 476.93 and a ClogP of 5.4 with extremely low solubility (<1 µg/ml).

Each formulation method was carefully screened: co-crystal screening was performed with

four solvents and 27 co-crystal formers, solid dispersion screening was conducted with

seven polymers and a total of 64 variants, and lipid-based formulations were screened with

25 GRAS listed excipients. Particle size reduction on a research compound from the same

scaffold at LCQ789 showed minimal impact on its oral bioavailability, as confirmed by

GastroPlus™ and in vivo rat study. Consequently, the nanoparticle approach was not

pursued for LCQ789.

Optimized formulations were compared in rat and dog single dose pharmacokinetic

studies. Minimal exposure was observed with suspensions of crystalline LCQ789 and co-

crystal formula in rats, but the co-crystal did show some improved exposure in dogs. In

both species, significant improvement in bioavailability was achieved with the solid

dispersion and lipid-based formulations. The in vivo pharmacokinetic studies indicated

that the solid dispersion improved the oral bioavailability by 18-fold in rats and 50-fold in

dogs, while the lipid-based formula increased the oral bioavailability by 25-fold in rats and

80-fold in dogs. The lipid-based formula achieved the highest exposure, and the solid

dispersion demonstrated less inter-subject variability. These two formulas continued on to

Page 32: Copyright by Chris Eugene Brough 2015

32

dose escalation studies. Higher exposure was achieved with the lipid-based formula, but

it also exhibited higher variability as compared to the solid dispersion. In addition, overall

high organic content was a concern for the long-term safety studies. As a result, the solid

dispersion formulation was selected for use in safety studies.

Vogt et al investigated micronization, co-grinding and spray-drying formulations

of fenofibrate and compared their release profiles to German and French commercial

products [74]. Micronization was conducted by jet milling while nanosizing was

performed by bead milling. Co-grinding was achieved by particle size reduction with

lactose, polyvinylpyrrolidone (PVP), sodium lauryl sulphate (SLS) and combinations of

the three excipients. Spray dried particles were produced by bead milling with lactose and

SLS in water, then fed directly into a Buchi Mini Spray Dryer. All formulations were

compared by dissolution in biorelevant media. The fenofibrate drug products

commercially available on the German and French markets dissolved similarly to crude or

micronized fenofibrate and were slower than the co-ground and spray-dried formulas. The

co-ground formulations had faster dissolution rates, but reached the same equilibrium

concentration as the commercially available products. The spray-dried formula produced

substantial initial supersaturation, but returned to concentrations only slightly better than

equilibrium in 180 minutes because formulation did not contain a parachute component.

This was termed as unstable by the authors, but they concluded that both co-grinding and

spray-drying could potentially lead to better bioavailability of fenofibrate drug products.

Kwong et al reported that toxicity is one of the leading causes of attrition in the

clinic, and that good safety margins are imperative to reach proof of concept in clinical

studies [75]. Research was conducted to identify a conventional formulation that would

provide the maximum exposure possible to define the dose limiting toxicity for a molecule

referred to as compound 3. Compound 3 existed as a crystalline free base with poor

Page 33: Copyright by Chris Eugene Brough 2015

33

solubility in water (0.0002 mg/ml in water, 0.0003 mg/ml in SGF and 0.004 mg/ml in

FaSSIF), but was highly permeable. Thus, drug dissolution was the rate limiting step for

absorption.

Figure 1.3: Dose proportionality of compound 3 in 10% Tween (-▲-) as the crystalline

form; in 0.5% Methocel/0.24% SLS as the amorphous form (-■-) and in

solid dispersion at 50% drug loading in HPMC-AS given as suspension in

Methocel (-●-) from 10 mpk to 750 mpk dosed at 5 ml/kg in Sprague

Dawley rats (n=4) [75] (corrected).

Various preparation methods and excipients were screened and the leading

formulations were compared in a preclinical dose range finding study in rats. The three

leading formulas were (1) crystalline form in 10% Tween 80, (2) amorphous form in 0.5%

methylcellulose and 0.24% sodium lauryl sulfate and (3) a solid dispersion with 50% drug

loading in HPMCAS-HF as seen in Figure 1.3. The particle size for compound 3 was not

specifically mentioned; however, looking at another reported project within the article, the

particle size is conservatively below 10 µm. At the lowest dose, 10 mpk, the exposure was

similar for all three compounds. As the dose was increased to 100 mpk, the solubility of

the crystalline phase limited the absorption of the compound and resulted in a lower

exposure. It is interesting to note that the exposure obtained from the amorphous form is

Page 34: Copyright by Chris Eugene Brough 2015

34

comparable to the solid dispersion at this level. Finally, at 750 mpk the amorphous

dispersion provided a significant increase in exposure over the other two formulations (4x

AUC as compared to the crystalline phase).

Thombre et al compared amorphous, nanocrystalline and crystalline formulations

of ziprasidone to commercially available Geodon® capsules in order to minimize food

effect [76]. The three formulations in the study were (A) an amorphous inclusion complex

of ziprasidone mesylate and a cyclodextrin, (B) a nanosuspension of crystalline ziprasidone

free base made by wet-milling, and (C) jet-milled ziprasidone HCl coated crystals made by

spray drying the drug with hypromellose acetate succinate. The pharmacokinetic studies

in dogs showed that the three formulations performed differently. Formulation B yielded

the best fasted state absorption enhancement, indicating that improved dissolution rate with

a low potential for precipitation might be the best combination to improve ziprasidone

absorption in the fasted state. The in vivo data for formulation B was highly variable, but

the mean summary parameters were comparable to the control capsule dosed in the fed

state.

Formulation A also showed increase absorption of ziprasidone in the fasted state

compared to the commercial capsule. Because of the solubilization technology used,

ziprasidone likely went quickly into solution and was absorbed (formulation A had the

shortest observed Tmax). But, the extent of absorption was less than the control capsule

dosed in the fed state, indicating possible precipitation in intestinal pH media. In vitro

testing demonstrated that some amorphous complex formulations can achieve high enough

supersaturation to cause precipitation of a lower solubility drug form such as the free base.

Formulation C did not show enhanced absorption in the fasted state compared to the

commercial capsule control.

Page 35: Copyright by Chris Eugene Brough 2015

35

In humans, both formulations A and B showed improved absorption in the fasted

state while formulation C did not [77]. Thus, the in vivo performance in dogs was

qualitatively similar to the in vivo performance of these formulations in humans.

Sigfridsson et al compared crystalline and amorphous nanosuspensions to a

solution of AZ68 [78]. AZ68 is a neurokinin NK receptor antagonist intended for

schizophrenia treatment. The compound has high permeability and low solubility in the

gastrointestinal track and thus fulfills the criteria for a BCS II compound. Crystalline

nanosuspensions were prepared by bead milling; amorphous nanosuspensions were

prepared by a solvent evaporation process and confirmed amorphous by XRPD (Powder

X-ray Diffraction). Particle sizes for both suspensions were approximately 200 nm and

both formulations were dosed in rats by gavage at similar concentrations. The results

indicate that AZ68 is absorbed at a lower rate for crystalline nanosuspensions compared to

amorphous nanosuspensions and solutions. However, the absorbed extent of the

compound is similar.

Page 36: Copyright by Chris Eugene Brough 2015

36

1.5 MARKETED PRODUCTS

1.5.1 Nano-crystal Products

Perhaps the best way to understand the utility of a technology is to review its

application in marketed products. Table 1.1 contains examples of current marketed

products produced using nano-crystal technologies.

Table 1.1: Examples of commercial nano crystal products [59]

The first commercial product using a pure nanoparticle technology was

Rapamune®, a tablet formulation containing the immunosuppressant drug, sirolimus. It

was originally marketed as an unpleasant tasting lipid based liquid solution that required

cold storage and a dispensing protocol. Sirolimus is poorly water soluble (~2 µg/ml) and

processed by NanoCrystal® technology into a fine nanoparticle dispersion with a mean

particle size of <200 nm. The NanoCrystal Colloidal Dispersion® intermediate is then post

processed into tablets of 1 mg, 2 mg and 5 mg. The tablets have increased bioavailability

(~ 27%), do not require special storage conditions and have better patient compliance due

to a more convenient dosage form [62, 79].

The second product, Emend®, is an antiemetic for the prevention of nausea and

vomiting following chemotherapy and surgery. It is a spray coated capsule formulation of

aprepitant that has been formulated as a nanosuspension. Aprepitant (water solubility 3-7

Trade Name Generic Name Processing Technology Company FDA Approval

Rapamune Sirolimus NanoCrystal Pfizer (Wyeth) 2000

Emend Aprepitant NanoCrystal Merck 2003

Tricor Fenofibrate NanoCrystal Fournier Pharma (AbbVie) 2004

Triglide Fenofibrate IDD-P Sciele Pharma, SkyePharma 2005

Megace ES Megestrole acetate NanoCrystal PAR Pharmaceuticals 2005

Invega Sustenna Paliperidone palmitate NanoCrystal Janssen 2009

Page 37: Copyright by Chris Eugene Brough 2015

37

µg/ml) is a free base crystalline compound that was processed into a fine particle dispersion

using a wet milling technology, and then processed into a solid dosage capsule formulation.

The original dosage form showed significant food effect with exposures of 3-fold increase

under fed conditions for a 100 mg dose. The food effect was significantly higher at a 300

mg dose [80]. The NanoCrystal® formulation was able to eliminate the food effect by

increasing surface area 40 fold which resulted in a 4 fold increase in AUC values in the

fasted state of beagle dogs. The elimination of the food effect for antiemetic drug is

significant as the commercial success may have been limited if required to consume with

food [62].

Tricor® is the successor product for fenofibrate (for hypercholesterolemia) after

patent expiration. Triglide is also a fenofibrate nano-crystal product, but produced by the

IDD-P® technology. Tricor® has a dose of 48 mg or 145 mg in a tablet form. The nano-

crystal technology provided a life-cycle extension in the creation of a superior performing

product. Fenofibrate showed 35% higher absorption in the fed state and significantly

reduced the food effect [32].

Megace ES® (ES stands for enhanced solubility) is for the delivery of megestrol

acetate, a synthetic progestin used to treat anorexia, cachexia and AIDS-related wasting.

The commercial formulation is an aqueous nanosuspension. The dose is 625 mg/ 5 ml and

the nanosuspension reduces bioavailability differences between fed and fasted conditions.

It also has less administration volume than the previously given oral formulation (only ¼)

while being less viscous [32].

The last approved nano-crystal product to be discussed is Janssen’s INVEGA®

SUSTENNA®. This product, an atypical antipsychotic, is a once monthly intramuscular

extended release injectable dosage form of paliperidone palmitate. It is a liquid dispersion

product in prefilled syringes that has physical stability without the need for special storage

Page 38: Copyright by Chris Eugene Brough 2015

38

conditions for a two year shelf life. This is a sterile product by way of conventional

sterilization methods. Terminal heat is the preferred route, but filtration, gamma irradiation

and aseptic approaches are employed as alternatives [81]. The particulate nature of the

formulation allowed for sustained release and in a patient population where compliance is

a concern, the once a month dosage is a valuable option [62].

1.5.2 Amorphous Dispersion Products

Table 1.2 contains examples of current marketed products produced using solid

amorphous dispersion technologies. Trade Name Generic Name Processing Technology Company FDA Approval

Cesamet Nabilone Solvent evaporation Meda Pharmaceuticals 1985

Sporanox Itraconazole Fluid bed bead layering Janssen 1992

Prograf Tacrolimus Spray drying Astellas Pharma 1994

Kaletra Lopinavir, ritonavir Melt extrusion AbbVie 2007

Intelence Etravirin Spray drying Janssen 2008

Zortress Everolimus Spray drying Novartis 2010

Norvir tablet Ritonavir Melt extrusion AbbVie 2010

Onmel Itraconazole Melt extrusion Merz Pharma 2010

Incivek Telaprevir Spray drying Vertex 2011

Zelboraf Vemurafenib Solvent/anti-solvent precipitation Roche 2011

Kalydeco Ivacaftor Spray drying Vertex 2012

Table 1.2: Examples of current amorphous dispersion products [82] (modified).

The original solid oral formulation of Kaletra® was a soft-gelatin capsule (SGC)

containing 133.3 mg of lopinavir (an HIV protease inhibitor) with 33.3 mg of ritonavir,

with ritonavir acting as a bioavailability enhancer for lopinavir [83]. The SGC dosage form

required refrigeration and the recommended adult dosage was 6 capsules daily with food

to maximize the bioavailability of lopinavir. Toward the end of end of enhancing therapy

Page 39: Copyright by Chris Eugene Brough 2015

39

with Kaletra, the product was reformulated as a solid amorphous dispersion utilizing HME

technology. The result was a 200/50 mg lopinavir/ritonavir tablet formulation that reduced

the number of dosage units and eliminated the requirement for refrigeration. Compared

with the SGC formulation, the HME formed tablet resulted in more consistent lopinavir

and ritonavir exposures across meal conditions, which minimized the likelihood of extreme

high or low blood plasma concentrations. In addition, the removal of the refrigeration

requirement allowed for worldwide distribution into underdeveloped areas that could not

support a cold distribution chain.

The path toward the development of ritonavir into a solid oral dosage form based

on amorphous solid dispersion technology was very similar to that of Kaletra. Ritonavir

was originally developed because it possesses anti-HIV activity, but it is no longer

prescribed as a sole protease inhibitor in antiretroviral regimens today. However, as a

pharmacokinetic enhancer, ritonavir has become a mainstay in the management of both

treatment-naïve and treatment-experienced patients. The basis for this enhancement is the

inhibition of cytochrome P-450 (CYP) metabolic pathways [84]. Inhibition of CYP 3A4

leads to a reduction in the metabolism of most protease inhibitors and increases

pharmacokinetic properties including area-under-the-plasma concentration curve (AUC),

maximum plasma concentration (Cmax), and half-life (t½). These enhancements allow for a

reduction of pill burden, dosing frequency and food restrictions while maintaining efficacy.

Because ritonavir is not bioavailable in the crystalline solid state, it was formulated

in ethanol/water based solutions. It was available as an oral solution in 1996 and as SGC

since 1998. Only one crystal form of ritonavir was identified during development and the

initial 240 lots of Norvir® capsules had no stability problems [85]. However, in mid-1998

several lots of capsules failed the dissolution requirement and investigation revealed a

polymorph with greatly reduced solubility compared to the original crystal form. Within

Page 40: Copyright by Chris Eugene Brough 2015

40

weeks, this new polymorph (Form II) appeared throughout both the bulk drug and

formulation areas. Since the manufacturing of Norvir® capsules required a certain

concentration in the ethanol/water solutions, Form II made the formulation unproducible.

With the emergence and dominance of polymorph Form II, Abbott Laboratories

reevaluated their manufacturing process [86]. They reported a new method to control the

conversion of Form I to Form II by choosing an appropriate ratio of solvent to antisolvent

in a reactor process. Following the HME formulation development for Kaletra® in 2005,

Abbott began developing a 100 mg HME-based amorphous dispersion formula for

ritonavir. In the bioequivalence studies, the HME tablet demonstrated equivalence with

SGC for AUC parameters. However, the ritonavir tablet Cmax was 26% higher than the

SGC formulation [84]. This increase in Cmax was not expected to alter the safety or

pharmacokinetic enhancing profile of ritonavir. The tablet also exhibits less food effect

than the SGC formulation.

Itraconazole is another interesting example of a drug product that was

commercialized using an amorphous solid dispersion technology. The compound is a

potent broad-spectrum triazole antifungal drug, is insoluble in water (solubility ~4 ng/ml)

and was among the first marketed solid amorphous dispersion products. Itraconazole is so

insoluble in intestinal fluids that drug therapy with the compound could not be achieved

without substantial solubility enhancement by formulation intervention. The original solid

oral formulation, Sporanox® Capsule, was produced by a fluid bed bead layering process

that used a co-solvent system of dichloromethane and methanol to dissolve itraconazole

and hydroxypropyl methylcellulose (HPMC) which was then sprayed on inert sugar

spheres [87]. The resultant product provided a significant enhancement of itraconazole

bioavailability with approximately 55% of the administered dose absorbed [88].

Itraconazole has recently been reformulated into a tablet composition that contains an

Page 41: Copyright by Chris Eugene Brough 2015

41

amorphous dispersion in HPMC 2910 by HME utilizing the MeltRx Technology®. The

trade name is Onmel®; it is available in a 200 mg strength for once-daily administration

and was approved by the FDA in April 2010 for the treatment of onychomycosis. The

HME formulation not only eliminated the use of organic solvents in manufacturing, but

also reduced dosing frequency from twice-daily to once daily [89].

Like itraconazole, Zelboraf™ is another compound for which the drug therapy was

enabled by the application of amorphous solid dispersion technology. The product is a

tablet dosage form containing an amorphous dispersion of vemurafenib in HPMCAS-LF

produced by a solvent/anti-solvent precipitation called microprecipitated bulk powder

(MBP) technology [90]. The process utilizes N,N-dimethylacetamide to dissolve the drug

and the ionic polymer. The solution is precipitated into acidified aqueous media, the

precipitates are then filtered, repeatedly washed to remove residual acid and solvent

content, dried, and then milled to form the amorphous powder intermediate, MBP. The

MBP product provides substantial enhancement of the solubility/ dissolution properties

and oral bioavailability of vemurafenib with excellent physical stability. The stability of

the amorphous dispersion is attributed to the high composite Tg, intermolecular interactions

between the drug and polymer as well as the moisture protective effect provided by the

polymer.

In initial Phase I clinical studies with a conventional formulation of vemurafenib,

patients did not respond, i.e. no tumor regression, to doses as high as 1,600 mg [91]. The

issue was identified as low oral bioavailability stemming from poor solubility, which

caused a halt to the clinical study until it could be reformulated into a more bioavailable

form. Due to melting point and organic solubility limitations, traditional amorphous solid

dispersion processes could not be applied, therefore necessitating the application of the

MBP technology. When clinical trials resumed with the new MBP-based formulation,

Page 42: Copyright by Chris Eugene Brough 2015

42

substantial tumor regression was achieved in a majority of patients as a result of the

enhanced formulation [92]. The application of the MBP technology to vemurafenib is a

compelling case study for the application of amorphous solid dispersion technology

because formulation intervention was directly responsible for enabling the drug therapy

and prolonging the lives of terminal patients suffering from metastatic melanoma.

Figure 1.4: Timeline of FDA approved amorphous dispersion and nano-crystal

products.

In concluding this section, viewing the nano-crystal and amorphous dispersion

FDA approved products on a timeline might indicate a possible trend in the drug

development industry (see Figure 1.4). From this analysis, it appears poorly water-soluble

drug candidates were formulated for commercialization in the 1980s and 1990s using solid

amorphous dispersion technologies. In the late 1990s it seems as if the industry turned to

nano-crystalline technologies as the method of choice for development of BCS class II and

IV active ingredients. However, from 2006 to 2012 only one nano-crystal product was

approved by the FDA while 8 amorphous solid dispersion products were accepted.

Page 43: Copyright by Chris Eugene Brough 2015

43

Additionally, the last two FDA approved nano-crystal products were suspensions; the last

one being an intramuscular injection product. This is a small sampling, but it looks like the

current industry preference for solving water-solubility challenges for oral delivery is

through the use of amorphous solid dispersion technologies.

Page 44: Copyright by Chris Eugene Brough 2015

44

1.6 CONCLUSION

Nano-crystal delivery forms and amorphous solid dispersions are well established

techniques for addressing poor water solubility in pharmaceutical compounds, however,

the methodologies are quite different. While several marketed products are made by both

nano-crystal technologies and amorphous solid dispersion technologies, it appears there

are more amorphous dispersion products on the U.S. market today. A timeline of FDA

approved drugs suggests current industry preference for solving solubility challenges for

oral delivery is leaning toward amorphous solid dispersion technologies.

Page 45: Copyright by Chris Eugene Brough 2015

45

Chapter 2

New Use of Polyvinyl Alcohol as a Solubility Enhancing Polymer for

Poorly Water-soluble Drug Delivery (Part 1)

2.1 INTRODUCTION

Poor water solubility has been documented as a common characteristic for new

chemical entities in development pipelines and commercially available products in the

pharmaceutical industry today [2-4]. Among the options for dealing with poor solubility,

the incorporation of the drug substance in a solid amorphous dispersion dosage form is

gaining popularity. There are many reasons for this increase in popularity with only a few

being mentioned here. First, conventional chemistry methods for solubility enhancement

like salt formation reportedly only work for 20-30% of new molecules, leaving the 70-80%

remaining to find other routes to improved solubility [11]. Second, solid dispersions

containing solubility enhancing polymers can dramatically increase apparent solubility for

durations sufficient to enable absorption form the intestinal lumen. Specifically, in the case

of ITZ, solubility increases of 10,000 fold or more have been demonstrated [40-42]. Third,

nano-crystal technologies have not proven as capable and industry trends show preference

to solid dispersions [1]. Fourth, compared with other noncrystalline drug delivery

approaches such as cosolvent systems or self-emulsifying drug delivery systems (SEDDS),

amorphous solid dispersions are more amenable to be developed into tablets, which is the

preferred solid dosage form for distribution, and amorphous dispersions demonstrate

highly desirable advantages over liquid or semisolid formulations including lower

manufacturing cost, smaller pill burden, and improved physical and chemical stability [57].

Finally, the various grades of polymers used for solid dispersions can not only increase

solubility, but also allow for targeted release and the tailoring of release profiles of a drug

substance. Thus, solving multiple problems simultaneously.

Page 46: Copyright by Chris Eugene Brough 2015

46

While numerous studies have been published on amorphous solid dispersions, there

are a limited number of polymers suitable for creating those systems. Polymers that have

been reported as a major component in solid dispersion formulations for improved

solubility are as follows: povidone (PVP) [93], copovidone (PVPVA) [94], crospovidone

(CrosPVP) [95], polyethyleneglycols (PEG) [96, 97], polymethacrylates [98, 99],

hydroxypropylmethylcellulose or hypromellose (HPMC) [87, 89], hypromellose acetate

succinate (HPMCAS) [100, 101], hypromellose phthalate (HPMCP) [102],

hydroxyethylcellulose (HEC) [103], hydroxypropylcellulose (HPC) [104], polyvinyl

acetate phthalate (PVAP) [42], cellulose acetate phthalate (CAP) [42], poloxamers [105],

carbomers [106], and Soluplus® [107]. As additional polymers are identified as options, a

broader range of challenging drug substances may be enabled for oral delivery systems.

Furthermore, existing prohibitive intellectual property could be circumvented and unique

release profiles could be created. Due to the high cost of development and regulatory

compliance verification, repurposing an existing excipient would seem a financial and time

efficient path.

Polyvinyl Alcohol (PVAL) is a water-soluble synthetic polymer represented by the

formula (C2H4O)n. The value of n for commercially available materials is between 500 and

5,000, which is roughly equivalent to a molecular weight range of 20,000 to 200,000 [108].

PVAL is unique among the vinyl polymers in the fact that the monomer, vinyl alcohol,

cannot exist in the free form. So, it is manufactured by the polymerization of vinyl acetate

and then converted by a hydrolysis (alcoholysis) process. Various grades based on extent

of hydrolysis exists with unconverted fractions being polyvinyl acetate. Since EMD

Millipore PVAL was used in this research, a brief explanation of their nomenclature is

prudent. Each grade of PVAL has two groups of numbers separated by a dash. The first

group of numbers represents the viscosity of the 4% aqueous solution at 20 °C as a relative

Page 47: Copyright by Chris Eugene Brough 2015

47

indication of the molar mass. The second group of numbers is the degree of hydrolysis of

the polyvinyl acetate. For example, an 88 in the second group would represent 88%

hydrolysis with 98% being considered a fully hydrolyzed grade. Thus, a 5-88 grade would

have the viscosity of 5 mPa·s in a 4% aqueous solution at 20 °C with the polymer chain

being approximately 88% PVAL and 12% polyvinyl acetate [109]. Table 2.1 contains

additional properties.

Available molecular weights 14,000 – 205,000 g/mol for partially hydrolyzed grades; 16,000 – 195,000 for

fully hydrolyzed grades.

Crystallite melting point (Tk) 180 – 240 °C

Glass transition temperature (Tg) 40 – 80 °C

Percent crystallinity Varies; increases with percentage hydrolysis.

Solubility in water Solutions can be prepared at any concentration. However, because viscosity

rises rapidly with solids content, certain practical limits exist for processable

solutions.

Solubility in organics Insoluble in most organic solvents. Slightly soluble in ethanol, practically

insoluble in acetone.

Temperature onset of degradation Approximately 180 °C

Aqueous solution pH (4% solution at room temperature) 5.0 – 6.5

Hygroscopicity DVS measurements reveal below a relative humidity of 50%, the change in

mass is less than 1%. At 90% relative humidity, mass changes of 8 – 17%

depending upon polymer type.

Table 2.1: Properties of PVAL (EMD Millipore) [109]

PVAL is currently being used in a variety of pharmaceutical applications. It is used

as a stabilizing agent for emulsions [110], in topical pharmaceutical [111] and ophthalmic

formulations [112, 113]. It is used in artificial tears [114] and integrated into contact lenses

for lubrication purposes [115]. Polyvinyl alcohol can be made into microspheres [116] and

is used as an emulsifier in creating PLGA nanoparticles [117, 118]. It has also been used

in sustained-release formulations for oral administration [119] and transdermal patches

[120]. Oral toxicity has been evaluated and found safe even at high levels of consumption;

Page 48: Copyright by Chris Eugene Brough 2015

48

the no-observed-adverse-effect level (NOAEL) for rats in a 3 month study was 5,000

mg/kg body weight / day [121].

To the authors’ knowledge, PVAL has not been reported as the polymeric carrier

in binary amorphous solid dispersions [122]. A brief discussion of the limitations of

conventional methods for creating amorphous dispersions provides insight to probable

reasons for this. Solvent processes (e.g., like spray drying and precipitation approaches)

require a drug/polymer solution prepared with a common solvent or co-solvents. PVAL is

insoluble in most organic solvents and only slightly soluble in ethanol while typical

pipeline drug molecules are only slightly soluble in more aggressive organic solvents.

Even if suitable solvents were ascertained, the required concentration of polymer would

render the solution too viscous for practical processing (like preventing flow through a

small orifice needed for spray drying). Alternatively, the application of conventional

fusion processes like hot melt extrusion (HME) and injection molding is limited due to the

high melting point of PVAL and high viscosity of the resulting melt. These technologies

are not suitable for processing thermally sensitive and viscous formulations, however

KinetiSol® has been demonstrated in these applications [123]. KinetiSol® Dispersing is a

novel fusion process that does not have torque limitations like HME and has been shown

to create unique solid dispersions [71]. Through this enabling technology, solid amorphous

dispersions containing PVAL as the primary polymer can now be evaluated.

The pH independence of PVAL presents some advantages in bioavailability

enhancement. The dissolution of pharmaceutical excipients with pH dependent ionizing

properties can be variable across the gastrointestinal pH range [124]. For example,

dissolution rate and apparent solubility of weakly basic drugs in gastric pH and weakly

acidic drugs in intestinal pH would be higher due to the ionization at those pH conditions

[40, 41, 125]. Anionic and cationic polymers are prone to those same ionization effects,

Page 49: Copyright by Chris Eugene Brough 2015

49

which allows for differing dissolution rates across the gastrointestinal tract. This may

become clinically significant for formulations utilizing ionic polymers in both inter-patient

and intra-patient variation. The dissolution behavior of non-ionic polymers is expected to

be the same across the gastrointestinal track, which should lower variability and provide a

more predictable bioavailability improvement [126]. Anionic polymers, even with

variation challenges, are still often chosen in formulation development due to their ability

to improve solubility compared to non-ionic polymers [127]. Anionic polymers have been

shown to increase dissolution rate and apparent solubility of both weakly basic and neutral

drugs in amorphous dispersions due to polymer/drug interactions likely occurring between

polymer functional groups and the drug substance [128]. The few non-ionic polymers

used for solid dispersions, primarily HPMC, PVP and PVPVA, do not exhibit tunable ratios

of hydrogen bond donors and receptors like HMPCAS and polymethacrylates. However,

PVAL has a tunable density of hydrogen bond donating hydroxyl groups and hydrogen

bond receptors along its polymer backbone while still having pH independent solubility.

In theory, this would allow it to function strongly as a solubility enhancer like the anionic

polymers, but also have the low variability of the non-ionic polymers.

It is hypothesized that the pH independent solubility along with tunable density of

functional groups of PVAL would be ideal in the creation of solid dispersions especially

for weakly basic drugs. While other groups have shown stability with semicrystalline

polymers, PVAL will need be investigated to establish the polymer crystalline structure

does not create nucleation sites for drug recrystallization during processing and storage

[129]. As a first step in investigating that PVAL will function as a solubility enhancing

polymer, ITZ was selected as the model drug to be formulated in various grades. Varying

molecular weight grades with identical degree of hydrolysis were used to create solid

dispersions to determine its effect on performance. Conversely, the best performing

Page 50: Copyright by Chris Eugene Brough 2015

50

molecular weight range of PVAL from the first group of formulations were selected in

grades of changing degrees of hydrolysis. The best performing polymer grade would then

be compared to commercially available ITZ products.

Page 51: Copyright by Chris Eugene Brough 2015

51

2.2 MATERIALS

Itraconazole was purchased from Hawkins, Inc. (Minneapolis, MN). All grades of

EMD Millipore PVAL were donated by Merck Millipore. All other chemicals used in this

study were ACS grade.

2.3 METHODS

2.3.1 KinetiSol® Dispersing (KSD)

All compositions for this study were produced by a lab scale, GMP pharmaceutical

compounder designed and manufactured my DisperSol Technologies, L.L.C.

(Georgetown, TX, USA). Prior to KSD processing, materials were weighed, placed into a

polyethylene bag, manually shaken for approximately 1 minute, and charged into the

compounder chamber. During processing, real time computer controls monitor various

parameters and eject the material at a pre-set ejection temperature. Discharged material

was immediately quenched in a cooling die under pressure in a pneumatic press. Cooled

material was then milled in a FitzMill L1A and screened through a 60 mesh screen (250

µm). All further analyses was conducted on this powder.

2.3.2 Differential Scanning Calorimetry (DSC)

Modulated DSC analysis was conducted using a TA Instruments Model mQ20 DSC

(New Castle, DE) equipped with a refrigerated cooling system and auto sampler. Samples

were weighed to 5 - 10 mg in aluminum-crimped pans. Samples were heated at a ramp rate

of 10 °C/min from 0 to 210 °C with a modulation temperature amplitude of 1.0 ºC and a

modulation period of 60 seconds for all studies. Ultrahigh purity nitrogen was used as the

purge gas at a flow rate of 50 mL/min. All data analyses were performed using TA

Universal Analysis software. The thermogram for amorphous ITZ used in the DSC analysis

Page 52: Copyright by Chris Eugene Brough 2015

52

of the solid dispersions formulations was obtained on a second heating of crystalline ITZ

following an initial heating to 200 ºC followed by rapid cooling (40 ºC/min) to 0 ºC.

2.3.3 Powder X-Ray Diffraction (XRD)

An Inel Equinox 100 X-ray diffractometer (INEL, Artenay, France) was used to

detect the presence of ITZ crystallinity. Milled compositions, physical mixtures or

unprocessed ITZ were loaded on a rotating aluminum sample holder and placed in the

radiation chamber. The Equinox 100 utilizes Cu K Alpha radiation (λ = 1.5418 Å) with a

curved radius detector to simultaneously measure a 2ϴ range of 5 - 110°. Operating voltage

and amperage were adjusted to 41 kV and 0.8 mA respectively and the scan time for each

sample was 10 minutes.

2.3.4 Polarized Light Microscopy (PLM)

While XRD has been shown to have limits of detection of less than 1% [130], PLM

can serve as a qualitative confirmation of XRD results. PLM analysis was conducted on a

Meiji Techno MT 9300 Polarizing microscope with a first order red compensator.

Pulverized samples were dusted on a glass slide and viewed under 400x magnification.

The slide holder was rotated at least 90 degrees while being observed to detect any light

refractions. Images were taken by an Infinity CMOS camera.

2.3.5 Size Exclusion Chromatography (SEC)

Molecular weights of PVAL in both processed and unprocessed samples were

analyzed with a Dionex Ultimate 3000 H/UPLC system equipped with a Shodex (Showa

Denko K.K., Japan) Refractive Index detector. Two Shodex Asahipak 7.6mm x 300mm

(GF-7M HQ) columns were placed series and maintained at 40 °C. A 50nM LiCl aqueous

solution was utilized at the mobile phase at a constant flow rate of 0.7 mL/min. Sample

injection volumes were 100 µL [131].

Page 53: Copyright by Chris Eugene Brough 2015

53

2.3.6 Non-sink Dissolution Analysis

Non-sink dissolution analysis was conducted with a VK 7000 dissolution tester

(Varian, Inc., Palo Alto, CA, USA) in accordance with USP XXXIV Method A for delayed

release dosage form. Solid dispersions (n=3) were weighed to achieve an equivalent of

37.5 mg ITZ and placed in dissolution vessels containing 750 ml of 0.1 N HCl (~10x ITZ

equilibrium solubility in acid) equilibrated to a temperature of 37.0 ± 1 °C with a paddle

rotation of 50 rpm. After 2 hours, a buffer medium of 250 mL of 0.2M Na3PO4, preheated

to 37.0 ± 1 °C, was added to the dissolution vessels to adjust the pH to 6.80 and reach 1000

mL in volume. Samples were taken at 1, 2, 2.25, 2.5, 3, 4 and 5 hours. Samples were

immediately filtered using 0.2 µm PTFE membrane, 13 mm filters, diluted to a 1:1 ratio

with mobile phase, mixed and then transferred into 1 mL vials for HPLC analysis.

2.3.7 High Performance Liquid Chromatography

The ITZ content in processed samples and dissolution aliquots were analyzed with

a Dionex Ultimate 3000 H/UPLC system equipped with diode array detector extracting at

263 nm. The system was operated under isocratic conditions with a 70:30:0.05

acetonitrile:water:diethanolamine mobile phase equipped with a Phenomenex Luna 5 µm

C18(2) 100 Å, 150 mm x 4.6 mm (Phenomenex, Torrance, CA) HPLC column. Dionex

Chromeleon 7.2 software was used to analyze all chromatography data.

2.3.8 In Vivo Studies

In vivo studies were conducted at Charles River Laboratories (Wilmington,

Massachusetts Facility) under NIH guidelines with IACUC approval. Male Sprague-

Dawley rats equipped with a surgically-implanted jugular vein catheters to facilitate blood

collection were prepared for the study. Animals were assigned to the study based on

catheter patency and acceptable health as determined by a staff veterinarian and placed into

Page 54: Copyright by Chris Eugene Brough 2015

54

groups of four animals (See Table 2.2 for details). All animals were fasted overnight prior

to dose administration and food was returned following the 4-hour post-dose blood

collection. All dosage forms were milled to a fine powder and suspended in an aqueous

dosing vehicle to insure accurate ITZ to rat body weight medicating.

Group

No.

No. of

Males

Treatment

Test Article Dose Level

(mg API/kg)

Dose Conc.

(mg API/mL)

Dose

Volume

(mL/kg)

Dose Vehicle Dose

Route

1 4 KSD

[ITZ:PVAL 4-88 (1:4)] 30 6 5 2% HPC/

0.1% Tween 80/

pH 2.0

Oral

gavage 2 4

Onmel™ tablets

(pulverized) 30 6 5

Table 2.2: Rat PK study design

Blood samples (0.25 mL; sodium heparin anticoagulant) were collected from the

jugular vein catheter or by venipuncture of a tail vein if the catheter became blocked. Blood

samples were collected from each animal at 2, 3, 3.5, 4, 5, 6, 8, 12, and 24 hours following

oral dosing. All whole blood samples were placed on wet ice immediately after collection

and were centrifuged at 2-8ºC to isolate plasma. The resulting plasma was transferred to

individual polypropylene tubes and immediately placed on dry ice until storage at

nominally -20C before transfer to the Charles River’s Bioanalytical Chemistry

Department for concentration analysis.

Plasma samples were analyzed for itraconazole concentration using a proprietary

in-house research grade LC-MS/MS Assay. Pharmacokinetic parameters were estimated

from the plasma concentration-time data using standard noncompartmental methods and

utilizing suitable analysis software (Watson 7.2 Bioanalytical LIMS, Thermo Electron

Corp).

Page 55: Copyright by Chris Eugene Brough 2015

55

2.4 RESULTS AND DISCUSSION

2.4.1 Preliminary Evaluation of PVAL Molecular Weight

Due to the uniqueness of the KinetiSol® technology, screening methods do not exist

to determine the processability of a new polymer. So, three viscosity grades representing

a low (PVAL 4-88), medium (PVAL 26-88) and high (PVAL 40-88) with the same level

of hydrolysis were selected with an arbitrary ITZ loading of 20%. Because of the feasibility

nature of this study, KSD processing parameters were not optimized but considered

acceptable if ITZ was rendered amorphous as determined by XRD and PLM analysis and

drug assays were at least 96%. This was easily achieved for the 4-88 grade as the

discharged material was a single agglomerated, semi-molten mass similar in character to

other previously processed formulations. The 26-88 and 40-88 grades behaved differently

in that the discharged material was a slightly densified powder. Variance in processing

RPM or increases in discharge temperature would not render an agglomerated discharge

which is typical of the KSD processes (preliminary results not reported). Other polymers

with larger molecular weights have been used in previous research without incident [132].

Thus, this phenomena must be attributed to the specific sintering properties or crystalline

nature of high viscosity grades of PVAL. The addition of 4-88 to 26-88 as a minor

component allowed for an agglomerated discharge, however a 1:1 ratio of 4-88 to 40-88

was required to achieve the same result. For consistency, a 1:1 ratio of 4-88 to both 26-88

and 40-88 was chosen as the compositions to continue this portion of the study.

Size exclusion chromatography was performed on unprocessed polymer grades as

controls and on the KSD processed compositions of ITZ with 4-88, 4-88:26-88 and 4-

88:40-88. Comparison of unprocessed and KSD processed 4-88 is shown in Figure 2.1A

while comparison of unprocessed, one-to-one physical mixture of 4-88 and 40-88 to the

KSD process composition is shown in Figure 2.1B.

Page 56: Copyright by Chris Eugene Brough 2015

56

A

B

Figure 2.1: SEC profiles of unprocessed powders and processed compositions.

Peaks associated with polymer have been numbered, followed by the elution time

(longer elution times signifies a smaller molecular weight). Each sample was analyzed in

triplicate, and representative chromatograms are presented herein. Unprocessed 4-88 had

a peak elution time of 24.917 minutes compared to the KSD processed 4-88 with a peak

time of 24.638. The physical mixture of 4-88 and 40-88 exhibited a first polymer peak at

21.193 minutes, which corresponds with the 40-88 grade, and the second peak at 24.788,

which is associated with the 4-88. The KSD composition showed elution times of the

respective peaks at 21.103 and 24.407. The small difference in the times can be explained

by sampling variance and it can be concluded that KSD processing did not change the

Page 57: Copyright by Chris Eugene Brough 2015

57

molecular weight of PVAL. However, it was noted that the KSD compositions always

eluded at an earlier time compared to the corresponding unprocessed powder, which would

indicate a slight increase in molecular weight. It is plausible that a portion of the

solubilized ITZ remains bound on the polymer following dissolution of the sample in the

diluent and while traversing the column, which would increase the apparent size of the

polymer coil and manifest as an increase in molecular weight. The key result of this

analysis is that there is no indication of a reduction in the molecular weight of the starting

PVAL material after KSD processing.

Dissolution profiles of the three compositions were compared (data not shown).

The ITZ:4-88 and ITZ:4-88:26-88 compositions performed similarly, but the ITZ:4-88:40-

88 composition did not maintain ITZ supersaturation to the same extent following the pH

change. With its ease of processing and good dissolution profile, the 4 mPa·s grade of

PVAL was selected as the molecular weight to continue evaluation of the effect of degree

of hydrolysis on polymer performance.

2.4.2 Solid-state Analysis

Various hydrolysis grades of the same viscosity were selected for comparison with

4-88, namely 4-38, 4-75 and 4-98. Polyvinyl alcohol is known to be a crystalline polymer

and XRD profiles were generated for each polymer used in this study to determine the

effect of molecular weight and degree of hydrolysis has on the crystallinity (see Figure

2.2). The XRD data matches the reported information in that crystallinity does appear to

increase with degree of hydrolysis. The 4-38 grade, contains a greater component of

polyvinyl acetate than PVAL, is the only grade that is XRD amorphous. All the other

grades had a primary peak between 19 and 20 2-theta with two other easily distinguishable

peaks between 22 and 23 and between 40 and 42 2-theta. The 26-88 and 40-88 grades

Page 58: Copyright by Chris Eugene Brough 2015

58

showed larger peak intensities than the 4-88 grade that might indicate increased

crystallinity with molecular weight and would explain why these grades were more

difficult to process at the selected conditions. The 4-98 grade, which is the only fully

hydrolyzed grade, demonstrated the highest peak intensities of all the grades tested.

Figure 2.2: XRD profiles of PVAL grades used in study

XRD profiles were also generated for ITZ and the processed compositions as

illustrated in Figure 2.3.

Figure 2.3: XRD profiles of ITZ and KSD compositions

Page 59: Copyright by Chris Eugene Brough 2015

59

For the 4-38, 4-75, 4-88 and 4-98 compositions, the profile demonstrates that the

drug substance is amorphous while the polymers retain their crystalline structure. In the

compositions with a one to one polymer ratio of low molecular weight with higher

molecular weight (only the 4-88:40-88 profile is shown, but the 4-88:26-88 profile was

identical), the major crystalline peak is visible, but with lower intensity. Other peaks are

barely distinguishable. It appears that the combination of different molecular weights

interact in such a way as to reduce the crystalline structure of the polymer portion of the

composite.

mDSC analysis was conducted on the pure PVAL polymers to understand the

thermal characteristic of each grade prior to analyzing the ITZ:PVAL KSD compositions.

Considering that PVAL is a copolymer of polyvinyl acetate (PVAC) and PVAL one would

expect to see by DSC analysis a Tg of the PVAC component and a melting endotherm of

the crystalline PVAL component, the prominence of both being a function of the degree of

hydrolysis. That is to say that with increasing degree of hydrolysis, the Tg of the PVAC

component will be less pronounced and the melting endotherm of PVAL will be more

distinct. Thus, a fully hydrolyzed grade would exhibit essentially no Tg associated with

PVAC and a strong melting endotherm associated with PVAL. The DSC thermograms for

the four 4 mPa·s grades of PVAL and their corresponding KSD compositions provided in

Figure 2.4 (nonreversing heat flow) and Figure 2.5 (reversing heat flow) demonstrate this

expected trend.

Page 60: Copyright by Chris Eugene Brough 2015

60

Figure 2.4: Nonreversing Heat Flow DSC thermograms for 4 mPa·s grades of PVAL

Figure 2.5: Reversing Heat Flow DSC thermograms for 4 mPa·s grades of PVAL

It is seen that the 4-38 grade, which is predominantly PVAC, exhibits a strong Tg

associated with the PVAC component and no detectable melting endotherm for PVAL

[133]. The amorphous nature of this grade as demonstrated by DSC is in agreement with

the previously discussed XRD result. Hence, it is understood that below a critical degree

of hydrolysis the PVAL component is unable to interact with itself to form crystallite

structures due to the density of PVAC on the polymer chain. Examining the thermograms

for the 4-75 and the 4-88 grades, the expected trend is seen in that the Tg becomes less

noticeable and the melting endotherm becomes more pronounced with an increased Tm as

Page 61: Copyright by Chris Eugene Brough 2015

61

the degree of hydrolysis is increased: consistent with increasing crystalline content of the

polymer. The thermal profile obtained for the 4-98 grade matches industry literature for

fully hydrolyzed polyvinyl alcohol with a melting point around 223 °C. The reversing heat

flow profile for PVAL 4-98 shows a slight thermal event around 70 °C which others have

suggested was the Tg [134]. This event was much less prominent than the Tg events for 4-

88, 4-75 and 4-38 as would be expected considering the low PVAC content and was not

able to be accurately quantified by the analysis software. The DSC profile for the 4-98

grade corroborates the XRD result indicating high crystallinity of the fully hydrolyzed

grade.

mDSC analysis was performed on the ITZ:PVAL KSD compositions to investigate

the dispersed state of ITZ in each polymer grade and to evaluate the influence of the

dispersed drug on the thermal properties of the polymers. Figure 2.6 contains the

nonreversing heat flow thermograms for the KSD composition and ITZ while Figure 2.7

contains the reversing heat flow thermograms.

Figure 2.6: Nonreversing Heat Flow DSC thermograms for KSD compositions and ITZ

Page 62: Copyright by Chris Eugene Brough 2015

62

Figure 2.7: Reversing Heat Flow DSC thermograms for KSD compositions and ITZ

The nonreversing heat flow thermogram in Figure 2.6 shows a general trend of

broadening and/or reduction in the magnitude of the glass transition event for the 4-38, 4-

75 and 4-88 grades relative to the pure polymers. This could be attributed to dilution and/or

anti-plasticizing effect of the dispersed ITZ phase. Consistent with the pure polymer, a Tg

is not apparent for the KSD composition with the 4-98 grade. Also seen for each of the 4-

38, 4-75 and 4-88 grades is an exothermic event indicative of ITZ crystallization with peak

temperatures in the range of 117 – 127 °C. Interestingly, the magnitude of the exothermic

event is inversely proportional to the degree of hydrolysis. This suggest that the physical

stability of an ITZ amorphous dispersion in PVAL increases with increasing degree of

hydrolysis. One possible explanation for this is that increasing the degree of hydrolysis

increases the density of hydrogen bon donor sites on the polymer backbone that could

interact with the hydrogen bond acceptor sites on ITZ to form a stable complex [86].

Another possible explanation is that the amphiphilic character of the polymer that stems

from the ratio of hydrophilic PVAL-to-hydrophobic PVAC units approaches an optimum

with regard to stabilizing interactions with ITZ as the percent hydrolysis is increased up to

88%.

Page 63: Copyright by Chris Eugene Brough 2015

63

When the degree of hydrolysis is low, as with the 4-38 grade, the dispersed ITZ

phase would interact largely with the PVAC component which contains only hydrogen

bond acceptor sites (assuming hydrogen bonding is forming the complex). Therefore, in

this phase, the amorphous dispersion is stabilized by weaker intermolecular forces like Van

der Waals forces. If the stable complex is formed by hydrophilic/hydrophobic

microenvironments, then the fewer hydroxyl groups of the 4-38 grade would provide fewer

areas for interaction. In either case, these weaker intermolecular forces are more easily

disrupted by thermal perturbations; hence, post-Tg phase separation of amorphous ITZ and

subsequent crystallization during the temperature ramp is observed to a greater extent with

reduced degree of hydrolysis [135]. This is corroborated by the result for the fully

hydrolyzed grade in which a crystallization exotherm was not detected.

Figure 2.6 also shows that for all KSD processed compositions, a minor ITZ

endotherm was detected. For the KSD compositions with PVAL 4-38, 4-75 and 4-88, this

endotherm is similar in magnitude to the crystallization exotherm, which suggest that the

ITZ melting event is related entirely to ITZ that crystallized during the DSC scan. This

result validates the findings of XRD in which no crystalline ITZ was detected. In the case

of the 4-98 grade, where no crystallization exotherm was detected, the ITZ melting

endotherm is likely due to residual crystalline ITZ remaining after processing. This could

be due to incomplete dispersion of ITZ in the polymer on processing resulting from the

highly crystalline nature and/or the high Tm of the PVAL 4-98 grade. It could also be the

result of amorphous ITZ being forced out of its interaction with the polymer as the highly

crystalline grade returned to its preferred crystalline orientation on quenching. This result

appears to contradict the XRD analysis; however, considering its small magnitude, the

melting endotherm represents a small fraction of the dispersed ITZ, the crystalline peaks

of which could be easily masked by the high intensity peaks of PVAL 4-98.

Page 64: Copyright by Chris Eugene Brough 2015

64

Finally, it is interesting to note in Figure 2.6 the melting endotherms for the PVAL

4-75 and 4-88 are significantly reduced in comparison to the pure polymer; beyond what

would be expected for a 20% dilution on the addition of ITZ. This suggests that the

dispersed ITZ phase disrupts crystallization of PVAL upon cooling due to drug/polymer

interactions. A similar melting endotherm, in terms of magnitude, between the pure

polymer and the KSD composition for the 4-98 grade seems to suggest incomplete

disruption of polymer crystallinity on processing and/or displacement of ITZ during the

quenching process. Considering this, it seems that a sufficient fraction of PVAC on the

polymer is required to disrupt PVAL crystallite structure to facilitate a molten transition

by KSD processing and achieve a homogenous dispersion of the drug substance [136].

Additionally, PVAC units interspersed on the polymer backbone may act as defects within

the PVAL crystallite structure that effectively make available sites for drug/polymer

interactions [97]. However, one would expect to see a point of diminishing returns for

increasing PVAC concentration as this would also lead to weaker drug/polymer

interactions. From Figure 2.6, it would appear that this optimum PVAL:PVAC ratio for

ITZ is near that of the 4-88 grade.

The reversing heat flow shown in Figure 2.7 shows similar trends to those of the

nonreversing heat flow analysis, however with better resolution of the glass transition

events. Specifically, the Tgs of the KSD compositions with the 4-88 and 4-98 grades are

able to be resolved by reversing heat flow analysis and have respective midpoint values of

58.05 °C and 51.78 °C. Also provided in Figure 2.7 is a thermogram for pure amorphous

ITZ which shows a Tg of 59 °C followed by two endothermic events at 74 °C and 90 °C

that have been identified as mesophases of glassy itraconazole [137]. In comparing all of

the KSD compositions in the figure, the 4-88 grade has the highest composite Tg. The Tg

is more than 10 °C higher (roughly 20%) than KSD compositions with 4-38 and 4-75

Page 65: Copyright by Chris Eugene Brough 2015

65

grades as well as the pure 4-88 polymer and is close to the Tg of pure amorphous

itraconazole. The anti-plasticizing effect of amorphous ITZ on the 4-88 grade indicates

strong positive drug-polymer interactions [138]; most likely due to hydrogen bonding

between the acceptor sites on ITZ and the donor sites on PVAL. The increased Tg relative

to the other PVAL grades seems to also support the previously discussed hypothesis

regarding an optimal ratio of PVAC to PVAL at which hydrogen bonding or

hydrophilic/hydrophobic locations are maximized. From these results, it can be concluded

that 4-88 is the best performer of the PVAL grades tested from a thermodynamic

standpoint.

2.4.3 Non-sink, Gastric Transfer Dissolution Analysis

As discussed previously, PVAL is an interesting new carrier polymer in the area of

supersaturating amorphous solid dispersions owing to its non-ionic nature and the high

density of hydrogen bond donor sites on the polymer chain. These attributes make it a

particularly attractive concentration enhancing carrier for weakly basic drugs which have

been shown to interact strongly with anionic carrier polymers to yield substantial

improvements in apparent solubility [40]. To investigate the concentration enhancing

effects of PVAL on ITZ, the dissolution properties of the four KSD compositions were

evaluated by a non-sink, gastric transfer dissolution method. The results of this analysis

are presented in Figures 2.8 and 2.9.

Page 66: Copyright by Chris Eugene Brough 2015

66

Figure 2.8: Dissolution Profiles for 4 mPa·s grades KSD compositions

Figure 2.9: Dissolution Profiles for 4 mPa·s grades KSD compositions presented as

concentration (C) relative to the saturation concentration (Cs) of ITZ at the

respective pH.

Both the 4-75 and 4-88 compositions achieved complete drug dissolution before

the pH change at 2 hours, while the 4-38 and 4-98 grades did not. Polyvinyl acetate is not

Page 67: Copyright by Chris Eugene Brough 2015

67

water soluble, which would explain why 4-38 did not go into solution like the higher

hydrolyzed grades. Owing to its highly crystalline nature, the 4-98 grade dissolves at a

very slow rate under the dissolution conditions of this study; hence, ITZ release from this

composition is slow and incomplete. After the pH change, the 4-88 grade provided

superior concentration enhancing effect presumably due to the greater PVAL content on

the polymer and increased probability of ITZ interaction with the polymer in solution. The

4-38 and the 4-98 grades provided essentially no concentration enhancing effect due to the

low concentration of dissolved polymer in the dissolution medium. Contrasting the

dissolution AUCs after the pH change, gives a rank order of 4-88 > 4-75 > 4-98 > 4-38

with respective values of 640 ± 89, 442 ± 23, 130 ± 9 and 115 ± 8 mg·min. By way of

comparison, DiNunzio et al. reported a post pH change dissolution AUC for Sporanox®

capsules of 226 ± 23 mg·min [42]. In Figure 2.9, these dissolution results are presented in

terms of concentration at each time point (C) relative to the saturation concentration (CS)

of ITZ at the respective pH. It should be noted that the CS for ITZ decreases dramatically

from 5 µg/ml to 1 ng/mL when the pH of the media is changed from 1.2 to 6.8 [5]. When

plotted in this manner, the extent of supersaturation enabled by PVAL 4-88 at neutral pH

becomes quite evident with solution concentrations nearly 9,000 times the saturation

concentration of ITZ. These results therefore demonstrate that PVAL is an effective

concentration enhancing polymer for ITZ and support the hypothesis that hydrogen bond

donor sites are the key points of interaction with weakly basic compounds that provide

stabilization/prolongation of aqueous supersaturation.

2.4.4 In Vivo Study

From solid state and dissolution evaluation of PVAL grades with varying degrees

of hydrolysis, PVAL 4-88 emerged as the top performing polymer. In order to investigate

Page 68: Copyright by Chris Eugene Brough 2015

68

the effect of this polymer on the oral absorption of ITZ from an amorphous dispersion, the

pharmacokinetics (PK) following oral administration to a rat model of this binary

composition were comparatively evaluated against the commercially available ITZ

product, Onmel™. The plasma concentration versus time profiles generated from this study

are shown in Figure 2.10 while the key pharmacokinetic parameters as calculated by non-

compartmental analysis are provided in Table 2.3 with the addition of previously obtained

results for Sporanox®. Table 2.3 contains the plasma drug concentration AUC(0-last), AUC(0-

∞ ), Cmax and Tmax.

Figure 2.10: Rat PK of PVAL 4-88 compositions compared with Onmel™.

Page 69: Copyright by Chris Eugene Brough 2015

69

AUC(0-last)

(hr·ng/mL)

AUC(0-∞)

(hr·ng/mL)

Cmax

(ng/mL)

Tmax

(hr)

ITZ:PVAL 4-88 (1:4) 3,315 ± 1336 3,595 ± 1662 391 ± 111 3.88 ± 0.25

Onmel™ 2,565 ± 1374 2,624 ± 1391 367 ± 202 3.50 ± 1.22

Sporanox® (From

[42])

2,132 ± 1273 359 ± 261 5.5 ± 2.3

Table 2.3: Rat PK mean calculations

The mean AUC and Cmax for the PVAL 4-88 composition was greater than that of

Onmel™ and Sporanox®. However, it is noted that the drug loading of the ITZ:HPMC

dispersion in the Onmel™ tablet is double that of the PVAL 4-88 composition. Similarly,

the nonpareil beads in the Sporanox® composition are coated with a 40% w/w itraconazole

and 60% w/w HPMC solid dispersion [139]. Therefore, the contrast with commercial

products is not a true apple-to-apple comparison. Further studies will evaluate PVAL

compositions at 40% w/w loading to have a more comparable result. Yet, considering the

exploratory nature of this study, the analogous in-vivo performance of this first generation

PVAL composition to commercial products is noteworthy.

Page 70: Copyright by Chris Eugene Brough 2015

70

2.5 CONCLUSION

This feasibility study has demonstrated that PVAL’s pH independent solubility

along with tunable density of functional groups allow it to function as a concentration

enhancing polymer, effectively increasing the apparent solubility of poorly water soluble

drug, itraconazole. The KinetiSol® technology enables the use of PVAL as the primary

carrier for amorphous solid dispersion compositions. Solid-state and non-sink dissolution

analysis revealed that the 88% hydrolyzed grade of PVAL was optimal for the ITZ

compositions. Pharmacokinetic analysis in a rat model demonstrated that the ITZ/PVAL

4-88 composition yielded exposures that were possibly superior to Sporanox® and

Onmel™.

Future research with ITZ will evaluate the effect of molecular weight in the range

between the 4-88 and 26-88 PVAL grades tested in this study on dissolution performance.

It will also explore increased drug loading levels, and evaluate storage stability. Additional

separate studies are planned to determine the broad application of PVAL evaluating

amorphous dispersions with neutral and weakly acidic drug substances. This

supplementary information would offer a foundation for development scientists, who are

looking for new options in formulating challenging drug substances, to have confidence in

exploring polyvinyl alcohol.

Page 71: Copyright by Chris Eugene Brough 2015

71

Chapter 3

New Use of Polyvinyl Alcohol as a Solubility Enhancing Polymer for

Poorly Water-soluble Drug Delivery (Part 2)

3.1 INTRODUCTION

One of the current challenges facing the pharmaceutical industry is addressing the

increasing presence of poorly water soluble drugs in commercial products and development

pipelines [2-4]. Using amorphous solid dispersions for oral delivery as a method to solve

these challenges has increased in popularity in both commercial products and industry

literature [1, 140-142]. For research purposes, itraconazole is often used as a model drug

for concentration enhancement due to its aqueous solubility of approximately 1 ng/ml at

neutral pH and approximately 4 µg/ml at 1.2 pH [41, 143, 144]. ITZ has been formulated

in various polymeric carriers including hydroxypropylmethylcellulose (HPMC) [87, 89],

copovidone (PVPVA) [94], polyethyleneglycols (PEG) [97], polymethacrylates [145],

hypromellose acetate succinate (HPMCAS) [146], hypromellose phthalate (HPMCP) [41],

Soluplus® [147], polyvinyl acetate phthalate (PVAP) and cellulose acetate phthalate (CAP)

[42].

The KinetiSol® Dispersing (KSD) technology is a thermal process than can create

solid amorphous systems from challenging drugs and very highly viscous polymers [71].

Utilizing this technology in our previous study, a broad range of molecular weights and

degrees of hydrolysis of polyvinyl alcohol (PVAL) were investigated as a primary carrier

in amorphous solid dispersion systems and for their solubility/bioavailability enhancing

effect on itraconazole. EMD Millipore’s PVAL was utilized in both studies and the

nomenclature for describing different polymer grades is X-Y. X represents the viscosity

(in mPa·s) of the 4% aqueous solution at 20 °C, which is a relative indication of the molar

mass and Y is the degree of hydrolysis of the polyvinyl acetate. An examination of the

Page 72: Copyright by Chris Eugene Brough 2015

72

effects of molecular weight were conducted using PVAL 4-88, 26-88 and 40-88 with the

4-88 grade performing the best both from a KinetiSol® processing and dissolution

enhancement point of view. Subsequently, 4-38, 4-75, 4-88 and 4-98 grades of PVAL

were tested to ascertain the effect of the degree of hydrolysis for dissolution performance

of ITZ. PVAL 4-38 did not fully release due to the fact that polyvinyl acetate, the

predominant component of the grade, is less water soluble than PVAL. PVAL 4-98 also

did not fully release due to the highly crystalline nature of the fully hydrolyzed grade which

swelled in the aqueous environment, but did not completely dissolve. Both PVAL 4-75

and 4-88 released fully in the acidic portion of the dissolution test, but after the pH change

to the neutral media the 4-88 grade provided best maintenance of supersaturated drug

concentrations.

Research conducted on various molecular weights of HPMC on the dissolution

performance of ITZ show that higher molecular weights are more efficient stabilizers of

drug supersaturation after the pH change from the acidic gastric condition to the neutral

intestinal environment [41, 132]. Similar to HPMC, PVAL has a wide spectrum of

molecular mass from low to high viscosity ranges and it is anticipated that it will have an

effect on the performance of the amorphous dispersion. Previous experimentation of

PVAL 26-88 and 40-88 showed that these grades are excessively viscous as to pose

processing challenges, which resulted in uneven drug distribution within the polymer.

However, there are several grades between 4-88 and 26-88, namely 5-88, 8-88, 18-88 and

23-88, which could be investigated to determine if molecular weight effects post pH change

dissolution performance. The current research comparatively evaluated the PVAL grades

within a processable viscosity range (4-88 to 18-88) for their effect on the properties of

amorphous dispersions of ITZ prepared by KSD.

Page 73: Copyright by Chris Eugene Brough 2015

73

Rendering a drug amorphous within a semi-crystalline polymeric carrier does raise

concerns for the long term storage stability of the formulation [148-150]. Since the drug

thermodynamically prefers the crystalline state, the polymeric carrier must prevent this

occurrence through steric hindrance or molecular interaction. The concern of semi-

crystalline polymers is that structured molecular geometry either allows for the formation

of polymer rich and drug rich domains or that the crystalline geometry negatively

influences the unstable amorphous drug. This is due to the stronger polymer-polymer

interactions with semi-crystalline polymers (compared to amorphous ones) and hence the

drug must compete with the polymer itself for binding sites on the formation and on the

storage of the dispersion. PVAL does have at least one advantage over other semi-

crystalline polymers like polyethylene glycol (PEG) in that it does have a higher Tg and

Tm, which results in lower molecular mobility at temperature ranges associated with

physical stability.

PVAL is known for providing solution stabilization as it is used in the

pharmaceutical industry both as a stabilizing agent for emulsions [110] and as an emulsifier

in creating PLGA nanoparticles [117, 118]. The previous research illustrated in both in

vivo and in vitro studies PVAL’s ability to increase free drug concentrations, and

consequently oral absorption of the small molecule, itraconazole. Because of this

performance, it is hypothesized that an in solution interaction exists between ITZ and

PVAL and can be characterized. In an attempt to determine the mechanism of interaction

that allows this apparent solubility increase, the current study employed the analytical tools

of Fourier transform infrared spectroscopy (FTIR) and dissolution-ordered spectroscopy

(DOSY).

Page 74: Copyright by Chris Eugene Brough 2015

74

3.2 MATERIALS

Itraconazole was purchased from Neuland Laboratories Limited (India). All grades

of PVAL (EMD Millipore) were donated by Merck Millipore. All other chemicals used in

this study were ACS grade

3.3 METHODS

3.3.1 KinetiSol® Dispersing (KSD)

Compositions for this study were produced by a lab scale, GMP pharmaceutical

compounder designed and manufactured by DisperSol Technologies, L.L.C. (Georgetown,

TX, USA). Prior to KSD processing, materials were weighed, dispensed into a

polyethylene bag, manually shaken for approximately 1 minute, and charged into the

compounder chamber. During processing, computer controls monitor processing

parameters in real time and eject the material at a pre-set ejection temperature. Discharged

material was immediately quenched in a cooling die under pressure in a pneumatic press.

Cooled material was then cryomilled in a SPEX 6870 Freezer/Mill, with a 2 cycle run of 5

minutes each at 10 cycles per second after a 3 minute cool time and 2 minute timer between

cycles. All further analyses was conducted on this powder.

3.3.2 Powder X-Ray Diffraction (XRD)

An Inel Equinox 100 X-ray diffractometer (INEL, Artenay, France) was used to

detect the presence of ITZ crystallinity. Milled compositions, physical mixtures and

unprocessed, pure ITZ were individually loaded on a rotating aluminum sample holder and

placed in the radiation chamber. The Equinox 100 utilizes Cu K Alpha radiation (λ =

1.5418 Å) with a curved radius detector to simultaneously measure a 2ϴ range of 5 - 110°.

Operating voltage and amperage were adjusted to 41 kV and 0.8 mA respectively and the

scan time for each sample was 10 minutes.

Page 75: Copyright by Chris Eugene Brough 2015

75

3.3.3 Polarized Light Microscopy (PLM)

PLM analysis was conducted on a Meiji Techno MT 9300 polarizing light

microscope with a first order red compensator. Pulverized samples were dusted on a glass

slide and viewed under 400x magnification. The slide holder was rotated at least 90

degrees while being observed to detect any light refractions.

3.3.4 High Performance Liquid Chromatography (HPLC)

The ITZ content in processed samples and dissolution aliquots were analyzed with

a Dionex Ultimate 3000 H/UPLC system equipped with diode array detector extracting at

263 nm. The system was operated under isocratic conditions with a 70:30:0.05

acetonitrile:water:diethanolamine mobile phase at a flow rate of 1.0 mL/min. The column

used for analysis was a Phenomenex Luna 5 µm C18(2) 100 Å, 150 mm x 4.6 mm

(Phenomenex, Torrance, CA, USA) HPLC column. Dionex Chromeleon 7.2 software was

used to analyze all chromatography data.

3.3.5 Non-Sink Dissolution Analysis

Non-sink dissolution analysis was conducted with a VK 7000 dissolution tester

(Varian, Inc., Palo Alto, CA, USA) configured as Apparatus 2. The test was performed

similar to USP XXXVII (38) dissolution for delayed release dosage forms. Solid

dispersions (n=3) were weighed to achieve a mass equivalent to 37.5 mg ITZ and dispensed

in dissolution vessels on the surface of 750 ml of 0.1 N HCl (~10x ITZ equilibrium

solubility in acid) equilibrated to a temperature of 37.0 ± 1 °C with a paddle rotation of 50

rpm. After 2 hours, a buffer medium of 250 mL of 0.2M Na3PO4, preheated to 37.0 ± 1

°C, was added to the dissolution vessels to adjust the pH to 6.80.

Drug content in solution was directly measured with a fiber optic Spectra™

instrument (Pion Inc., Billerica, MA, USA) fitted with 5 mm path length Pion Probes.

Page 76: Copyright by Chris Eugene Brough 2015

76

Readings were recorded every 5 minutes over the entire duration. Additionally, samples

were taken at 2:00, 2:05, 2:10, 2:15, 2:30, 3:00, 3:30, 4:00 and 5:00 hours. Samples were

immediately filtered using 0.2 µm PVDF with GMF membrane, 13 mm filters, diluted at a

1:1 ratio with mobile phase, mixed and then transferred into 1.5 mL vials for HPLC

analysis. Pion data was correlated to match HPLC results by selecting appropriate 2nd

derivative parameters.

3.3.6 Precipitation Inhibition Analysis

Solid dispersions (n=3) were weighed to achieve an equivalent of 37.5 mg ITZ and

placed in suitable vessels containing 750 ml of 0.1 N HCl (to match concentrations in non-

sink dissolution studies) that had been equilibrated to a temperature of 37.0 ± 1 °C.

Containers were placed on stirring tables and agitated vigorously for several hours until

solutions became clear. Media was then transferred to dissolution vessels within a VK

7000 dissolution tester (Varian, Inc., Palo Alto, CA, USA) configured as Apparatus 2 with

a paddle rotation of 50 rpm. A buffer medium of 250 mL of 0.2M Na3PO4, preheated to

37.0 ± 1 °C, was added to the dissolution vessels to adjust the pH to 6.80. Samples were

taken immediately before pH change, then after the pH change at 5, 10, 15, 30, 60, 90, 120,

and 180 minutes. Samples were immediately filtered using 0.2 µm PVDF with GMF

membrane, 13 mm filters, diluted to a 1:1 ratio with mobile phase, mixed and then

transferred into 1.5 mL vials for HPLC analysis.

3.3.7 Fourier Transform Infrared Spectroscopy (FTIR)

Samples were tested in a Thermo Scientific Nicolet iS50 FT-IR Spectrometer

(Thermo Electron Scientific Instruments LLC, Madison, WI, USA) with the Omni-

Sampler module for attenuated total reflectance (ATR). The scan range was 2000 – 700

Page 77: Copyright by Chris Eugene Brough 2015

77

wavenumber (cm-1) at a resolution of 4 cm-1 for 32 scans per sample. Data collection and

analysis were performed with OMNIC™.

3.3.8 Diffusion-Ordered Spectroscopy (DOSY)

Amorphous dispersions and PVAL were dissolved in deuterated solutions of 0.1N

DCl prepared by diluting concentrated DCL with D2O. Samples were tested in a Varian

VNMRS 600MHz nuclear magnetic resonance (NMR) spectrometer (Palo Alto, CA, USA)

for solution with a Doneshot pulse sequence. The spectra was adjusted until water at 25

°C was 19.02 x 10-10 and all data generated is correlated to that reference. Figures were

generated using Varian VNMRJ software.

Page 78: Copyright by Chris Eugene Brough 2015

78

3.4 RESULTS AND DISCUSSION

3.4.1 Evaluation of PVAL Molecular Weight

To be consistent with our previous study, a drug loading of 20% ITZ was selected

for the preparation of amorphous dispersions with PVAL grades of 4-88, 5-88, 8-88 and

18-88. All acceptable KSD compositions were verified to be amorphous by XRD and

PLM, and HPLC analysis was conducted to establish dispersion potencies (data not

shown). For dissolution testing, the samples were weight adjusted according to potency to

achieve an equivalent amount of ITZ in each vessel. Figure 3.1 contains the dissolution

curves of the listed PVAL grades for the 2 hour pH change dissolution method. Error bars

are not shown due to data density making them indistinguishable.

Figure 3.1: Dissolution profiles of 20% ITZ in PVAL 4-88, 5-88, 8-88 and 18-88.

Both PVAL 4-88 and 5-88 grades fully released ITZ at supersaturated levels with

little variance between the two in the acid phase. PVAL 8-88 and 18-88 were not able to

Page 79: Copyright by Chris Eugene Brough 2015

79

achieve complete release, which was observed during the test with powder appearing at the

bottom of the dissolution vessels after approximately an hour. After the pH change the

precipitation inhibition performance followed the reverse order of molecular weights: 4-88

> 5-88 > 8-88 > 18-88.

Prior to the pH change, 8-88 had only reached 70% of drug in solution and 18-88

had only attained 60%. It was speculated that if these two grades could achieve 100% drug

release the rank order of performance might change. An alternative dissolution study using

the same amorphous dispersions, called precipitation inhibition analysis, was conducted

which forced complete drug release from all polymer grades prior to the pH change with

the results exhibited in Figure 3.2.

Figure 3.2: Precipitation inhibition curves of 20% ITZ in PVAL 4-88, 5-88, 8-88 and

18-88.

Even with the formulations at the same starting point, the lower molecular weights

outperformed the larger ones in reducing the rate of drug precipitation. These results match

Page 80: Copyright by Chris Eugene Brough 2015

80

the previous study where PVAL 4-88 provided better performance than combinations

containing the 26-88 and 40-88 grades. Thus, contrary to the trends observed with HPMC,

for PVAL the lower molecular weights have superior dissolution performance over higher

molecular weights.

This trend is confirmed by comparing the rates of dissolution for the polymers

themselves in water. Product literature shows that lower molecular weights of the -88

grades have faster dissolution rates specifically in the temperature range between 20 and

40 °C [109]. Additionally, as shown in that same product literature, PVAL reduces

interfacial tension, particularly the surface tension of water with respect to air, with

differing effects according to molecular weight as seen in Figure 24. That figure

specifically compares PVAL grades 40-88, 18-88, 8-88 and 4-88 for reduction in water

surface tension. The rank order of greatest surface tension reduction to lowest is 4-88 > 8-

88 > 18-88 > 40-88. Thus, the lower molecular weights have a greater reduction in surface

tension of water.

Since amorphous dispersion powders were dispensed on the surface of the media

in the vessels for the non-sink dissolution test, the lower surface tension and faster

dissolution rate of the lower molecular weight grades would speak to the faster dissolution

rate of ITZ. The effect on water surface tension indicates that lower molecular weight

grades effectively function as a polymeric surfactant more than higher molecular weights.

This would support the results for the precipitation inhibition analysis where the lower

molecular weights were more effective in reducing the rate of drug precipitation.

3.4.2 Drug Loading Evaluation

All dissolution studies from the current and previous research have been conducted

with 20% drug loading. It is noted that the drug loading of ITZ:HPMC in the commercial

Page 81: Copyright by Chris Eugene Brough 2015

81

product Onmel™ and the coating for the nonpareil beads in Sporanox® which is also ITZ

in HPMC are both 40% [139]. Typically, the higher the drug load, the lower the dissolution

performance because the ratio of drug to polymer decreases and solubility enhancement

trends with increasing polymer concentration [151]. However, if higher drug loading can

be achieved, the size of the dosage form can be reduced, typically improving patient

convenience and compliance. Hence, optimal drug loading is a balance between

performance and size of the final dosage form.

To investigate the effect of drug loading on dissolution performance, amorphous

dispersions of ITZ in PVAL 4-88 were produced at 10, 20, 30, 40 and 50% drug loads.

HPLC analysis measured actual potencies and samples were weight adjusted to ensure 37.5

mg of drug were contained in each dissolution vessel. The result from the non-sink

dissolution test is found in Figure 3.3.

Figure 3.3: Dissolution comparison of different ITZ loadings in PVAL 4-88.

Page 82: Copyright by Chris Eugene Brough 2015

82

The rank order from highest dissolution performance in acid is as follows: 20% >

30% > 40% > 10% > 50%. Surprisingly, the 10% dispersion did not have the highest rate

of dissolution. It exhibited a faster initial rate than 30% and 40%, but slowed after the

initial 20 minutes to perform only better than the 50% loading. Only the 10% and 50%

drug loading did not achieve 100% drug release within the two hour window of the acid

portion of the experiment.

Toward the end of the 2 hour acid phase of the dissolution test, powder was visible

at the bottom of the vessel of the 50% composition. All other vessels were clear at the

bottom, however there was still powder floating at the top of the vessel of the 10% drug

load sample. Keeping the relative amounts in perspective can provide probable

explanations for this behavior. The 10% drug loading has a theoretical weight of 375 mg,

the 20% drug loading has 187.5 mg weight on down to the 50% drug loading only having

75 mg of total amorphous intermediate powder. The 187.5 mg of 20% ITZ loaded powder

floated on top (like all other samples) initially, then within a few minutes dispersed into

the vessel and was dissolved within 20 minutes. Both the 30% and 40% loaded samples

followed the same pattern, but with longer timeframes. In the case of the 10% ITZ

dispersion, the time for the floating material to disperse within the aqueous environment

and then dissolve was so prolonged that a very small amount of powder remained on the

top surface at the end of the 2 hour acid phase. Visually, it was only a fraction of the initial

375 mg amount but it is unclear as to whether it could account for all of the undissolved

amount of ITZ.

Results from the previous study and product literature show that increasing polymer

crystallinity reduces polymer solubility and dissolution rate. The lowest drug loading at

10% would have the highest allowance for the formation, in some regions, for the preferred

crystalline configuration which would slow dissolution and negatively affect solubility.

Page 83: Copyright by Chris Eugene Brough 2015

83

With the more crystalline configuration, ITZ could be entombed in highly crystalline

regions which would account for the incomplete release and the more sustained drug

solubility after the pH change due to higher polymer concentrations. Increasing drug load

would decrease polymer crystallinity and allow for the faster dissolution for the 20%

sample. However, increasing the drug loading also increases the hydrophobicity of the

dispersion resulting in slower wetting and dissolution as seen with the 30% and 40%. The

50% dispersion would have the lowest polymer crystallization, but also the most reduced

wetting and diminished release rate due to low polymer concentration.

After the pH change the rank order follows as was expected with 10% > 20% >

30% > 40% > 50%. 10% ITZ did start at a lower concentration than the 20%, but had

higher actual amounts of drug in solution at 180 minutes which was an hour after the pH

change. The 20% ITZ composition had the fastest dissolution rate and reduced the rate of

precipitation at a similar level to the 10% composition. The 30% loading was a close

second to 20% in the acid phase dissolution and exhibited a rate of precipitation only

slightly lower than both the 10% and 20% compositions. Since, optimal drug loading is a

balance between performance and size of the final dosage form, the 30% loading is

probably closer to the optimal drug loading than all the other drug loadings tested.

3.4.3 Evaluation of the ITZ/PVAL Interaction

Because the ITZ is amorphous in the solid dispersion, it is not surprising to see

apparent solubility levels much higher than the crystalline equilibrium solubility.

However, without the aid of a solubility enhancing/stabilizing polymer, the

thermodynamically unstable supersaturation would soon return to the crystalline

equilibrium solubility. Especially, as the aqueous environment transitions from the acid

gastric environment to the pH neutral environment of the intestinal tract, a weakly basic

Page 84: Copyright by Chris Eugene Brough 2015

84

drug like ITZ would return to solubility equilibrium quite rapidly [132]. Dissolution

profiles contained in this and the previous study report concentration levels as high as 8,000

times the saturation concentration in neutral media and maintaining at least 4,000 times the

saturation concentration for 30 minutes after the pH change. These results suggest an

attractive interaction between the ITZ and PVAL.

FTIR-ATR has been reported as an analytical tool to investigate such possible

interactions [53, 152, 153]. Peaks in the spectra can be associated with specific atomic

bonding and shifting or changing shape can indicate molecular interactions. Figure 3.4

contains the FTIR spectra of amorphous itraconazole, preprocessed PVAL 4-88 and

amorphous solid dispersions of 10% and 20% ITZ loading in PVAL 4-88.

Figure 3.4: FTIR-ATR spectra of amorphous ITZ, PVAL 4-88, and amorphous

dispersions.

Page 85: Copyright by Chris Eugene Brough 2015

85

The result for ITZ is similar to published literature [154, 155] and the PVAL result

is consistent with product literature [109]. For PVAL, the broad peak at 3330 cm-1 is

associated with O-H stretching, the small peak at 2941 cm-1 with C-H stretching, the 1732

cm-1 peak with C=O groups (from unhydrolyzed ester groups), the peak at 1429 cm-1 with

CH2 bending, the 1251 cm-1 peak with the C-O-C groups (also from unhydrolyzed ester

groups) and, finally, the 1093 cm-1 peak is associated with C-O stretch and O-H bending.

For ITZ, the first major peak at 1700 cm-1 is associated with C=O stretching, the 1511 cm-

1 peak is related with CO-NH2, and the 946 cm-1 linked with C-Cl. Most peaks do not show

any difference between the individual components and the amorphous dispersions. Yet,

the broad peak associated with O-H stretching on PVAL showed a shift from 3330 cm-1 to

3334 and 3335 cm-1 on the 10% and 20% ITZ loading in PVAL 4-88 respectively. The

ITZ peak at 1700 cm-1 that correlates to C=O stretching, shows a change of 1713 and 1706

cm-1 for the same respective 10% and 20% ITZ loaded amorphous dispersions. As

mentioned previously, the 10% amorphous dispersion will exhibit a higher degree of

polymer crystal formation than the 20% load, which might explain the larger shift for the

lower drug load. While it does appear that hydrogen bonding might be occurring, the

magnitude of the shifts advocates that interactions, more than just hydrogen bonding alone,

would account for the storage stability of the amorphous dispersion.

Diffusion-ordered spectroscopy (DOSY) or sometimes called diffusion-ordered

NMR spectroscopy and diffusion-ordered correlation spectroscopy has been used to

determine the drug complexation with cyclodextrins [156], drug release from a hydrogel

matrix [157], identification of fake formulations of commercial products [158], and to see

the interactions of polymers and drugs in solution [159]. DOSY is a measure of diffusion

and can distinguish different components in a formulation by the rate of diffusion and

location on the spectra. Small molecules have a faster diffusion rate than large polymers

Page 86: Copyright by Chris Eugene Brough 2015

86

and will have a different measurement of diffusion unless they are strongly bound by a

covalent, ionic or hydrogen bond, which would give them the same measurement of

diffusion as they act as single entity. Figure 3.5 has the DOSY map for PVAL 4-88 only

and for the 20% drug loading in PVAL4-88.

Figure 3.5: DOSY maps of (A) PVAL 4-88 and (B) 20% ITZ in PVAL 4-88 amorphous

dispersion.

The Y or F1 axis is the measurement of diffusion or how well the molecules travel

within the solution with lower numbers at the top representing slower diffusion and

increasing with the numbers to reach the highest diffusion rates at the bottom. The X or

F2 axis helps track any chemical shift from the top spectra. Figure 3.5A is a map of pure

PVAL 4-88 that has not been processed. It is interesting to note that the polymer has two

different diffusion coefficients, with the first band around 0.4 and the second one around

A B

Page 87: Copyright by Chris Eugene Brough 2015

87

1.0 on the F1 axis. Product literature speaks specifically about PVAL’s propensity toward

complex formation in the presence of certain acids or salts [109]. Since the solution is

0.1N DCl, it is possible to have PVAL form a bimolecular complex. Additionally, in the

discussion regarding PVAL reduction of water surface tension, it was mentioned that

PVAL 4-88’s greater effect on surface tension would allow it to function with surfactant-

like properties, which would allow for molecules to self-associate in solution. Thus, it

appears that the two different diffusion coefficients is formed by some molecules going

into solution in groups and other by themselves. Whatever the actual cause is for the

grouping phenomena, it appears that acetate groups on the PVAL molecular chain play a

part because the spectra associated with acetate (around 2 on the F2 axis) is only observed

in the slower diffusion group. Also on Figure 3.5A is an area that is around 10 on the F1

axis and around 2 on the F2 axis; this charts the residual acetyl groups within the bulk

polymer.

Figure 3.5B is a map of the amorphous dispersion containing 20% ITZ in PVAL4-

88. At the top, again, is the two different diffusion coefficients of the PVAL 4-88. Around

3 on the F1 axis is the spectra of itraconazole, which illustrates a different measurement of

diffusion for the drug than the polymer. This illustrates clearly that the drug is not tightly

bound to the polymer by strong intermolecular interactions. If there is any hydrogen

bonding, the drug/polymer association would have to be flexible enough to allow ITZ

movement along PVAL to account for the different rates of diffusion. Ideally, a third

DOSY map of only ITZ in solution would be very useful to see if the interaction of the

polymer changes the diffusion for the drug to illustrate a practical interaction between the

two, but, unfortunately, the concentration of ITZ in solution was below the limit of

detection for this analytical test. There is an area around 2 on the F1 axis that lies between

the diffusion of PVAL 4-88 and ITZ that could possibly be a slower measured diffusion

Page 88: Copyright by Chris Eugene Brough 2015

88

for ITZ that would indicate an interaction. However, it could also be a software

interpretation error due to peaks of both PVAL 4-88 and ITZ both occurring in that area.

Thus, the key conclusion that can be drawn from these results is that the concentration

enhancing effect of PVAL on ITZ is likely the result of weak intermolecular interactions,

such as van der Waals forces/hydrophobic interactions, as it appears that strong

intermolecular interactions between ITZ and PVAL do not occur in solution. This

conclusion is supported by other research efforts that have looked at amorphous dispersions

of ITZ with various polymeric carriers and concluded that hydrogen bonding was not the

mechanism of action in solution between the drug and polymer [42, 132].

3.4.4 Evaluation of Stability

One of the main concerns to be addressed when creating amorphous dispersions is

the stability of the matrix to prevent the drug from returning to its crystalline state within

the delivery form before dosage [54, 57, 160]. Physical stability of the drug substance can

be achieved through kinetic stabilization or freezing the amorphous drug within the

polymeric carrier. This can be achieved by having a Tg of the composite matrix 50 °C

above ambient storage conditions [52]. It can also be achieved by thermodynamic stability

which is attained by molecular interactions that provide stability [56]. Many amorphous

dispersions are formulated in a metastable region where the mode of stabilization is a

combination of both kinetic and thermodynamic mechanisms [57]. Drug loading also

influences stability because as loading is increased, at some point miscibility is exceeded

and then drug molecule proximity within the polymer matrix becomes critical. Stability

tends to decrease because recrystallization becomes more probable as drug molecules are

increasingly found in close proximately to other drug molecules.

Page 89: Copyright by Chris Eugene Brough 2015

89

The drug loading portion of the current study concluded that 30% ITZ in PVAL 4-

88 is probably closer to the optimal drug loading than all the other drug loadings tested. A

container (without desiccant) containing milled amorphous intermediate powder of 30%

ITZ was stored at ambient conditions for 30 months. A new sample that was identical in

formulation was prepared and compared to the 30-month-old material by XRD and PLM.

The XRD diffractogram is shown in Figure 3.6.

Figure 3.6: XRD profiles of ITZ, initial and 30 month old amorphous dispersions.

Both the initial and 30 month XRD profiles contain no peaks that correspond with

ITZ indicating that both are substantially amorphous. The major broad peak in the

amorphous dispersions at approximately 19 2-theta and minor peaks at approximately 23

and 41 2-theta are related to the crystallinity of the PVAL 4-88. No visual crystal structures

were observed in either samples by PLM. Since it is possible to have crystals below the

limit of detection of XRD and PLM, a dissolution study was conducted to confirm the

dissolution performance of the amorphous compositions. These results are presented in

Figure 3.7. As stated in the methods section, the first 120 minutes of the dissolution was

Page 90: Copyright by Chris Eugene Brough 2015

90

performed in 0.1 N HCL after which a buffer is added to change to a neutral pH. Thus, in

the acidic environment full release of ITZ is realized after which the pH change resulted in

precipitation of ITZ. Since only two dissolution profiles are compared in this figure, error

bars were included to show the significance of any variation.

Figure 3.7: pH change dissolution comparing initial and 30 month ambient storage

amorphous dispersions of ITZ:PVAL 4-88 (30% drug loading).

The dissolution performance in 0.1 N HCl of the two compositions are comparable

with the 30 month old composition performing slightly better. Two different lots of PVAL

4-88 were used to make these compositions, which could explain the slight variance.

Another possibility is that the stored material had some water absorption that would allow

for better wetting and faster release. Since the purpose of the test was to determine crystal

growth during storage, the reason for this variance was not investigated. After the pH

change, the two compositions performed similarly with the two curves staying within the

error bars. Because the dissolution performance is similar, it can be concluded that the

Page 91: Copyright by Chris Eugene Brough 2015

91

solid-state properties of ITZ in the PVAL matrix did not change during storage. Therefore,

dissolution performance and XRD results demonstrate acceptable real-time ambient

stability for the lead ITZ:PVAL 4-88 composition.

Page 92: Copyright by Chris Eugene Brough 2015

92

3.5 CONCLUSION

PVAL has been further investigated as a solubility enhancing polymer in

amorphous solid dispersions as enabled by KinetiSol® Dispersing. PVAL 4-88 was

confirmed to provide the best solubility enhancement and in solution stability as compared

with 5-88, 8-88 and 18-88 presenting a trend that lower molecular weights function better

than higher molecular weight polyvinyl alcohols. In a drug loading study, 20, 30 and 40%

ITZ drug loads achieve full drug release in 0.1 N HCl in 2 hours, and were able to retard

the return to the crystalline equilibrium solubility after the pH change in the expected rank

order of 10%>20%>30%>40%>50% with the 30% loading selected as optimal. DOSY

and FTIR were used confirm that weak molecular interactions are the likely cause of solid

state and solution stability of amorphous ITZ. A 30 month stability study showed the

optimal ITZ:PVAL 4-88 amorphous dispersion remained entirely amorphous as

determined by XRD and PLM and the dissolution performance was unchanged.

Page 93: Copyright by Chris Eugene Brough 2015

93

Chapter 4

Novel formulation approach for multiparticulate abuse deterrent

development

4.1 INTRODUCTION

The misuse and abuse of prescription drugs poses a serious public threat in the

United States today. The Centers for Disease Control and Prevention (CDC) have declared

that the United States is in the midst of an epidemic of prescription drug overdose deaths

[161]. The CDC reported that drug overdose deaths were second only to motor vehicle

crash deaths among the leading causes of unintentional injury death in 2007 [162]. That

same report issued statistics that the number of deaths involving opioid analgesics was 1.93

times the number involving cocaine and 5.38 times the number involving heroin. A 2010

Substance Abuse and Mental Health Services Administration survey found that 16 million

Americans age 12 years and older had taken a prescription pain reliever, tranquilizer,

stimulant, or sedative for nonmedical purposes at least once in the previous year; 7 million

had used psychotherapeutic drugs nonmedically within the past month. Of these drug

abusers, 55% said they obtained the drug they most recently used from a friend or relative

for free [163]. These statistics express the rising need for drug formulators and regulators

to create barriers to prescription drug abuse.

To combat this growing epidemic, the U.S. Food and Drug Administration has

released a guidance document for drug formulation companies to follow for formulating

prescription drugs prone to abuse [164]. That guidance document offers several suggested

approaches that are summarized in Table 4.1.

Page 94: Copyright by Chris Eugene Brough 2015

94

Approach Brief Description

Physical/chemical

barriers

Physical barriers can prevent chewing, crushing, cutting,

grating or grinding of the dosage form. Chemical barriers

can resist extraction of the drug using common solvents like

water or organic solvents.

Agonist/antagonist

combinations

A drug antagonist can be added to interfere with, reduce or

defeat the euphoria associated with abuse.

Aversion Substances can be added to the product to produce an

unpleasant effect if the dosage form is manipulated or is used

at a higher dose than directed.

Delivery system Certain drug release designs or method of delivery can offer

resistance to abuse like a sustained-release subcutaneous

implant.

New molecular entities

and Prodrugs

The properties of a NME or prodrug could include the need

for enzymatic activation, different receptor binding profiles,

slower central nervous system penetration or other effects.

Combination Two or more of the above methods could be combined to

deter abuse.

Novel approaches This category encompasses novel approaches that are not

captured in the previous categories.

Table 4.1: Categories for abuse deterrence as defined by the U.S. FDA.

The first category, physical and chemical barriers, represent the most prevalent

approach in currently available commercial products. The most financially successful

commercial abuse deterrent formulation has been Purdue Pharma’s OxyContin®, which

Page 95: Copyright by Chris Eugene Brough 2015

95

was reformulated in a polyethylene oxide matrix as a physical barrier and approved by the

FDA in April 2010 [165, 166]. The original product label did not contain any information

on abuse deterrence despite having in vitro evidence to support such a claim. After product

launch several studies investigated the public health impact of the OxyContin®

reformulation to support the relabeling of the product with abuse deterrent information.

One study looked at patients entering treatment programs with opioid dependency [167].

The selection of OxyContin® as a primary drug of abuse decreased from 35.6% of

respondents before the release of the abuse-deterrent formulation to just 12.8% 21 months

later. Those same respondents reported that of all opioids used to ‘get high in the past 30

days’ OxyContin® fell about 40% while the use of heroin nearly doubled over the same

time period. Another study reported that abuse of the reformulated OxyContin® was 41%

lower than historic abuse of the original formulation [168]. ‘Real world’ in vitro testing

demonstrated the relative difficulty of manipulation of the reformulation [169] along with

human in vivo testing [170] assessed the actual abuse deterrent properties of the

reformulation. While not completely tamper proof, the reformation of OxyContin® did

prove to be a deterrent in all evaluations.

The use of a polyethylene oxide (PEO) matrix has been utilized in Opana® (an

oxymorphone reformulation by Endo Pharmaceuticals) and in Nucynta® (a tapentadol

reformulation by Ortho-McNiell-Janssen Pharmaceuticals) but neither have been allowed

to state abuse deterrence on product labels by the FDA even with some post-marketing

studies [166]. Remoxy® (Pain Therapeutics, Inc.) is an oxycodone extended release

product that does not use PEO to provide a physical barrier but rather a formulation

containing sucrose acetate isobutyrate in a gelatin capsule [171]. The formulation creates

a viscous mass that resists size reduction even when frozen and only releases a fraction of

the oxycodone when submerged in high-proof alcohol for hours. Clinical trials showed

Page 96: Copyright by Chris Eugene Brough 2015

96

significant pain reduction and survey data from drug counselors demonstrated that it was

less preferred for abuse as compared to the original ER oxycodone formulation [172, 173].

However, the FDA issued a complete response letter in June, 2011, indicating additional

studies were required before approval would be granted. Since sucrose acetate isobutyrate

is not considered GRAS by the FDA, there appears to be an opportunity for creating a solid

matrix product out of GRAS excipients that does not contain PEO as a barrier approach in

abuse deterrent formulation.

There are commercial products that belong to other abuse deterrent categories.

Suboxone® is categorized as an agonist/antagonist combination of buprenorphine and

naloxone, where the buprenorphine acts as a partial opioid agonist combined with the

opioid agonist naloxone. It was approved in 2003 by the FDA for the treatment of opioid

dependence but is commonly used off-label as a pain therapeutic [174]. Embeda® is also

an agonist/antagonist combination of morphine and naltrexone. When taken orally as

intended, the naltrexone is not released. But, if the tablet is chewed or crushed prior to

taken orally, the naltrexone is released with the morphine and the agonist activity is limited

[175]. This complex formulation encountered manufacturing issues and was recalled in

March, 2011. Oxecta® is an immediate release formulation of oxycodone that contains

aversive ingredients. Since the active ingredient is readily available via the oral route, the

aversion abuse deterrent approach is to reduce non-oral abuse. The formulation contains

sodium lauryl sulfate which irritates the nasal cavity to deter insufflation and it also

contains excipients that form a gel when exposed to water to deter intravenous use [176].

Dysphagia, or difficulty in swallowing tablets, greatly influences patient

compliance. A study has shown that up to ¼ of the general population may have problem

swallowing, with the very young having the most difficulty followed by elderly patients

[177]. Unfortunately, these age groups are commonly issued prescriptions for drugs that

Page 97: Copyright by Chris Eugene Brough 2015

97

are likely to be abused. For example, in 2007, 3.1 million youths aged 12 to 17 received

treatment or counseling for problems with behavior or emotions with a significant

percentage receiving psychotherapeutic drugs [178]. On the other hand, different studies

have shown that the age group most likely to suffer from chronic pain is age 50 to 64 [179,

180]. It would be highly beneficial to have a delivery form that could simultaneously

provide ease of dosing for patent compliance while at the same time delivering abuse

deterrence.

One such approach would be to have a physical barrier formulation in a

multiparticulate delivery form rather than a single, larger abuse deterrent tablet. The

smaller particles would allow for oral delivery by sprinkling onto soft food for consumption

to help with patient who suffer with dysphagia. In addition to ease of dosing, there are

other biopharmaceutical advantages with having a multiparticulate delivery form.

Gastrointestinal tract (GIT) transit time for small particles would not depend upon gastric

emptying, which would result in a more uniform distribution in the GIT and less variable

pharmacokinetic profiles among patients. However, the challenge with a multiparticulate

approach is that the smaller particle size has higher susceptibility to certain mechanical

abuse techniques such as size reduction by a coffee grinder or mortar and pestle. The high

surface area also lends faster drug release when submerged in ethanol, which is also a

common abuse technique. So, while the theory of a multiparticulate delivery is appealing,

the technical details to the approach are challenging.

Currently, the most published multiparticulate approach is a platform technology

called DETERx (Collegium Pharmaceuticals, Inc.) that creates a complex of drug with

fatty acid and then places it with a hydrophobic waxy matrix in submillimeter, spherical

particles for extended release [181-183]. The spherical particles are prepared using a

standard spray congealing process and are specifically designed to retain the time-release

Page 98: Copyright by Chris Eugene Brough 2015

98

mechanism following common methods of tampering (e.g. crushing, chewing, heating for

IV injection, etc.). The release profile can be adjusted by excipient ratios and particle size.

Not much information is available on the challenge of creating the drug/fatty acid ionic

complex, but a simple drug-in-a-matrix approach would be less complex and would

possibly allow for a broader range of drugs to be incorporated in the platform.

The KinetiSol® Dispersing (KSD) technology is a thermal process than can create

solid amorphous systems from challenging drugs and very highly viscous polymers [71].

While most publications on this technology deal with amorphous systems, the process can

produce dispersions of crystalline drug just as straightforward as amorphous ones.

Polymers with viscosity too large for process via extrusion have been successfully

processed with KSD with the resulting solid systems being very difficult to mill once

cooled [132]. While there is more to abuse deterrent systems than being difficult to mill,

the viscous polymers do inherently provide resistance to crushing and powder formation.

Additionally, the flexibility of KSD for polymer selection will allow for formulating with

polymers that have inherent ethanol resistance like using polyvinyl alcohol (PVAL). It is

hypothesized that through KSD, very viscous polymers like high molecular weight

hydroxypropylmethylcellulose (HPMC) can be combined with other excipients to create

abuse deterrent formulations. Furthermore, if combined with excipients that retard drug

release, very viscous polymers can be formed into a multiparticulate extended release

delivery form, providing dosing benefit to patients with dysphagia while not needing the

complex processing of DETERx.

Page 99: Copyright by Chris Eugene Brough 2015

99

4.2 MATERIALS

Theophylline (THEO) was purchased from Acros Organics (Thermo Fisher

Scientific). HPMC grades were received from Dow Chemical Company or purchased from

Ashland Chemical Company. Ethocel (EC) grades were also sourced from Dow Chemical

Company. TEC and DBS was purchased from Vertellus. EVAC and PVAL grades were

purchased from Sigma-Aldrich.

4.3 METHODS

4.3.1 KinetiSol® Dispersing (KSD)

Compositions for this study were produced by a lab scale, GMP pharmaceutical

compounder designed and manufactured by DisperSol Technologies, L.L.C. (Georgetown,

TX, USA). Prior to KSD processing, materials were weighed, dispensed into a

polyethylene bag, manually blended for approximately 1 minute, and charged into the

compounder chamber. For compositions with liquid components, the non-liquid

components were weighed and placed into a food processor. The liquid components were

then weighed out in syringes that had been tared after being flushed with the liquid

component (to compensate for the weight of residual liquid in the syringe after injection).

The liquid was slowly added to the non-liquid components and then mixed on low setting

for 2 minutes. During processing in the compounder, computer controls monitor

processing parameters in real time and eject the material at a pre-set ejection temperature.

Discharged material was immediately quenched in a cooling die under pressure in a

pneumatic press. Cooled material was then milled in a FitzMill L1A with a 0.156 inch

(approximately 5 mesh) screen after which particles smaller than a 12 mesh screen size

were removed from the resulting material. Formulations with high impact resistance

required cooling with liquid nitrogen prior to milling.

Page 100: Copyright by Chris Eugene Brough 2015

100

4.3.2 Non-Sink Dissolution Analysis

Non-sink dissolution analysis was conducted with a VK 7000 dissolution tester

(Varian, Inc., Palo Alto, CA, USA) configured as Apparatus 2. Samples equivalent to 37.5

mg THEO were weighed and added to the surface dissolution vessels filled with 750 ml of

0.1 N HCl equilibrated to a temperature of 37.0 ± 1 °C with a paddle rotation of 50 rpm.

Drug content in solution was directly measured at 270 nm with a fiber optic Spectra™

instrument (Pion Inc., Billerica, MA, USA) fitted with 5 mm path length Pion Probes.

Readings were recorded every 10 minutes for 24 hours. For dissolution containing ethanol,

300 ml of ethanol was added to 450 ml of 0.1 N HCl and equilibrated to a temperature of

37.0 ± 1 °C prior to adding samples.

4.3.3 Dart Impact Test

Dart impact tests were conducted on a custom made testing device. Essentially, a

dart (solid steel cylinder) weighing 1.198 kg is placed with in PVC tube that is mounted

vertically. A string and pulley system allows the dart to be raised to a specific height and

then released. At the base of the PVC tube is a flat steel plate where a tablet is placed to

be hit by the falling dart. The dart is raised to 213 cm in height so that the tablet is hit with

25 J of energy. A tablet passed the test if it remained intact in one piece without visual

cracks.

4.3.4 Coffee Grinder Size Reduction Test

Approximately one gram of the multiparticulates is weighed out and placed in a

KitchenAid coffee grinder. The coffee grinder is turned on for 15 seconds followed by a

15 second pause. This is repeated for a total time of 5 minutes. The ground material is

then placed on top of a screen stack containing 20, 40, 60, and 200 mesh with a pan at the

bottom to collect all fines that would pass through the 200 mesh. The screen stack was

Page 101: Copyright by Chris Eugene Brough 2015

101

mechanically shaken for two minutes. Each screen was weighted prior to sieving and then

weighted afterward to determine the total weight of material in each size range.

4.3.5 Syringeability Test

Approximately 500 mg of the multiparticulates were placed on a metal pan and held

over a propane torch set on the lowest setting. The height from the pan to the flame was

slowly reduced until the heat was sufficient to start melting the particle surfaces. Once

molten, the pan could be tipped so the melted material could pool on the lip of the pan and

be drawn up by a 20 gauge needle. The needle and syringe were pre-weighed so that any

material drawn up by the needled could be quantified by weighing the needle, syringe and

drawn material.

4.3.6 Water Extraction Test

Three 150 ml beakers were filled with 100 ml of distilled water and placed in three

different conditions. The first beaker was left at room temperature, the second one was

placed on a hot plate and brought to a mild boil, and the last one was placed in an ice water

bath to reach approximately 3 °C. Once the three beakers achieved the desired conditions,

25 mg of multiparticulate were placed in the water. Drug content in solution was directly

measured at 270 nm with a fiber optic Spectra™ instrument (Pion Inc., Billerica, MA,

USA) fitted with 5 mm path length Pion Probes. Readings were recorded every 2 minutes

for 3 hours. Test was concluded as soon as the water in the heated beaker reached a level

that would not allow for complete submersion of the fiber optic tip. The final volume was

measured and concentrations were calculated on the average of the beginning and final

volumes.

Page 102: Copyright by Chris Eugene Brough 2015

102

4.4 RESULTS AND DISCUSSION

4.4.1 Formulation Development

In the previous chapters and in other research [132] where amorphous solid

dispersions were created using high molecular weight polymers, it was discovered that

milling the KSD discharged material was a challenge. It was also discovered if the drug

was not rendered amorphous, then the discharged material was significantly easier to mill.

In general, it is known that drugs can function as plasticizers once they are rendered

amorphous or molten and it has been found that ITZ works well as a plasticizer in cellulosic

polymers [184]. Since drugs that are abused typically do not need bioavailability

enhancement, it would be preferred to make solid dispersion with drugs in their unaltered

or crystalline state. However, not only would the composition not gain the benefit of the

drug acting like a plasticizer, but the small, crystalline molecules would also have a

negative impact on mechanical properties of the polymer.

To compensate for this, a small component of the formulation must have some

plasticizing effect to retain impact properties sufficient to resist crushing or milling. A

screening study was conducted with 15 different excipients with HPMC E50 and 20%

THEO by processing the compositions by KSD and then directly forming a tablet for dart

impact testing. For confidentiality reasons, the 15 excipients will not be listed in this

chapter. Six of the formulations successfully passed the impact test; two of them were

selected to move forward with formulation development: triethyl citrate (TEC) and dibutyl

sebacate (DBS). TEC was selected because of the elastomeric properties that it imparted

on HPMC, which makes it particularly impact resistant. DBS was selected because it is

less water soluble than TEC and might provide a slower release in dissolution testing.

The desired release rate from an extended release delivery form is drug and therapy

dependent. In this study, THEO is being used as a surrogate for an opioid like oxycodone

Page 103: Copyright by Chris Eugene Brough 2015

103

because of its high melting point and water/alcohol solubility. For pain therapy, opioids

have been tested for in 4-6 hour and 12 hour release [185]. For a starting point, the authors

thought that a release profile around 6 hours could be representative of a target product

profile and was selected as the development goal. Several formulations were made from

the selected group of materials, HPMC, EC, PVAL, EVAC, TEC and DBS, and tested in

dissolution. Formulations that required at least 3 hours to achieve full drug in solution are

listed in Table 4.2. Along with the time to achieve full release, the table also contains an

impact rating, which was a subjective categorization after manually bending the discharged

after cooling in the compression mold. A rating of 1 was very easy to break, a rating of 2

required some effort to break, and a rating of 3 was very difficult or impossible to break.

Formulation Time to full

release

Impact

rating

62% HPMC E50 / 10% HPMC K100M / 8% TEC 3:00 3

44% PVAL 4-38 / 36% K100M 3:10 2

40% PVAL 4-88 / 40% PVAL 4-98 3:20 1

40% EVAC 12 / 30% HPMC K100M / 10% EC 10 4:00 1

40% EVAC 12 / 40% HPMC K100M 4:10 2

52% HPMC E50 / 10% EC 10 / 10% HPMC K100M / 8% TEC 4:10 2

42% HPMC E50 / 20% EC 10 / 10% HPMC K100M / 8% TEC 4:30 1

30% HPMC E15 / 30% HPMC K100M / 20% DBS 5:40 3

35% HPMC E15 / 35% HPMC K100M / 10% DBS 6:40 3

55% EC 10 / 25% DBS 35% in 24:00 1

80% EVAC 12 20% in 24:00 3

Table 4.2: Formulations tested to determine excipient trends (n=1).

Many of the formulations initially tested attained full release in around 2 hours, so

the formulation focus turned to determining combinations that released THEO very slowly.

EVAC is used for extended release transdermal drug delivery [186] and is used in the

plastic industry as an impact modifier [187]; consequently, it was not surprising that it

Page 104: Copyright by Chris Eugene Brough 2015

104

required more than 24 hours to release and received an impact rating of 3. EVAC was

introduced into formulations containing HPMC E50 and it did slow the release. But, the

time to full release was still faster than three hours and the impact rating of the

compositions were 2 or less. When combined with HPMC K100M, a release time of 4

hours was achieved but again the impact rating was 2 or less. Ethyl Cellulose is used in

formulations for extended release and lends itself well to thermal processing [188]. Mixed

with a high percentage of DBS yielded a composition that required more than 24 hours to

release. However, EC is brittle and very easy to grind or mill. It was added at 10 and 20%

in several formulations and generally aided in slow down release. Unfortunately, the

negative impact on the mechanical properties even at 10% was sufficient to make all

formulations a rating of 2 or less even with the impact augmenting effects of TEC.

HPMC K100M in combination with both plasticizers proved to be the best

mechanical property enhancing approach in formulation development, but most of the

release times were between 3 and 4 hours. Other HPMC grades were added to the

formulations both to make the composition easier to process and to alter the release profile.

Eventually, a combination of E15, K100M and DBS generated a release profile around 6

hours and earned an impact rating of 3.

As mentioned in the introduction section, one of the common methods for abuse is

to consume alcohol simultaneously with the dosage form to defeat the extended release

feature and attain dose dumping. Thus, the last step of the formulation process was to see

the release profile in 40% ethanol compared with the release in 0.1 N HCl to determine if

further formulation adjustments were needed. Figure 4.1 contains the dissolution profiles

of the two conditions for the lead formulation of 20% THEO / 35% HPMC E15 / 35%

HPMC K100M / 10% DBS.

Page 105: Copyright by Chris Eugene Brough 2015

105

Figure 4.1: Dissolution profiles of lead formulation in 0.1 N HCl and 40% Ethanol

(n=3).

The lead formulation actually had a slightly slower dissolution rate in the media

containing 40% ethanol than it did the standard media. An F2 similarity factor was

calculated to compare the two dissolution profiles and a value of 90 was obtained [189].

This indicated that there is less than 2% variance between the two profiles. Thus, the lead

formulation meets to the main performance criteria of releasing the drug over 6 hours,

having good mechanical properties for impact resistance, and having a release profile that

is not significantly altered by ethanol.

4.4.2 Abuse Testing

A reduction in particle size will cause an increase in the dissolution rate. Thus, one

of the abuse methods is to reduce the delivery form to a fine powder to have an immediate

0.0

20.0

40.0

60.0

80.0

100.0

0:00 1:30 3:00 4:30 6:00 7:30

% D

rug

Rel

ease

Time

0.1 N HCl Media

40% Ethanol Media

Page 106: Copyright by Chris Eugene Brough 2015

106

release rate. The multiparticulates produced in this study are in the size range of 1.7 to 4

mm. All the particles that were smaller than a 12 mesh screen were removed from the

multiparticulates, so the smallest size corresponds with the 12 mesh screen which is 1.7

mm. The 4 mm size on the large end of the spectrum corresponds with the milling screen

size. The coffee grinder size reduction test was conducted on the lead formulation with the

results of the mass fraction analysis shown in Figure 4.2.

Figure 4.2: Post Coffee Grinder Test Particle Size distribution.

The small sieve stack was limited in sizes, so the percentage of particles larger than

12 mesh could not be measured. However, almost 62% of the multiparticulates were over

the 20 mesh size which shows that very little size reduction did occur on the majority of

the particles. None of the particles were reduced in size to a fine powder (less than 75

microns) which is a positive result to keep the delivery form from being rendered an

immediate release. The remaining particles were scattered in sizes ranging from 75 to 850

0.0

10.0

20.0

30.0

40.0

50.0

60.0

70.0

<75 75-250 250-425 425-850 850+

0.0

12.5 14.810.9

61.8

Mas

s Fr

acti

on

(w

t%)

Particle Size (µm)

Page 107: Copyright by Chris Eugene Brough 2015

107

microns. The real indication of the effect of the coffee grinder size reduction test is to

perform a dissolution comparison which is illustrated in Figure 4.3.

Figure 4.3: Dissolution Profile of Original Size and Coffee Ground Material (n=3).

As expected, the ground material had a faster dissolution rate than the original size

material, but it was not an immediate release profile. The T95 (time where 95% of the

drug has been released) and T80 (time for 80%) was 4:42 and 2:32 for the original size and

3:32 and 1:42 for the coffee ground material, which is approximately a time reduction of

30%. An F2 similarity factor was calculated for the entire profile and for the first three

hours with values of 84 and 67 respectively, which is a difference of roughly 5%. While

it is a change in the dissolution performance, the ground material still retained an extended

release profile to serve as an abuse deterrent.

0.0

20.0

40.0

60.0

80.0

100.0

0:00 1:30 3:00 4:30 6:00 7:30

% D

rug

Rel

ease

Time

Original Size

Ground Material

Page 108: Copyright by Chris Eugene Brough 2015

108

Skeptics of this test might argue that the coffee grinder was only run intermittently

for 5 minutes and if it were processed consistently for a longer duration of time that the

particle size reduction would allow for immediate release. Prior to having access to a

Fitzmill in early KSD research studies, coffee grinders were used for size reduction of the

discharge of amorphous dispersions. Several coffee grinders were broken until a method

was discovered to grind high molecular weight polymers without causing damage to the

grinder. If the coffee grinders were run continuously, the surfaces of the particles are

heated up by the friction occurring in the grinder. As the surface temperatures of the

particles rises, they become soft and tacky. The tacky surfaces create a much higher

resistance to rotation, which will result in breaking the drive shaft or burning out the electric

motor. It is likely that an individual attempting to abuse the delivery form for the first time

would not take such precautions and cause damage to the coffee grinder, which would

probably motivate the person to change to a different method of tampering.

A mortar and pestle was employed to see if size reduction could be achieved.

Figure 4.4 is a picture of the material after test to show that no particle size reduction or

powder was formed.

Figure 4.4: Mortar and Pestle approach for size reduction of multiparticulates.

Page 109: Copyright by Chris Eugene Brough 2015

109

After 5 minutes of vigorously attempting size reduction in the mortar and pestle, no

size reduction was observed. A hammer test was briefly attempted, where the

multiparticulates were placed in a polyethylene bag and hit with a hammer. The first

impact did substantial damage to the plastic bag and the second impact caused some

multiparticulates to fly across the room. No size reduction was observed and the test was

ceased for safety reasons. With the results from the coffee grinder, mortar and pestle, and

hammer testing, it was concluded that the lead formulation did provide good resistance to

size reduction.

The syringeability test was then conducted on the lead formulation as illustrated in

Figure 4.5.

Figure 4.5: Result of Syringeability test of lead formulation.

As the particles were heated up, the surfaces in contact with the pan would melt

and then instantly char. Some of the particles in the figure demonstrate this as the bottom

is black while the top is still unmelted. It is possible that the heat transfer rate is too high

from the propane torch and causing the surface to burn before the top portion of the particle

Page 110: Copyright by Chris Eugene Brough 2015

110

received enough heat to melt. Consequently, the test was conducted on a hot plate where

the temperature was slowly raised over approximately 15 minutes to allow time for

sufficient heat transfer. Even with the slower heat transfer rate, the particles burned on the

bottom and remained solid on the top (picture not shown). Thus, this action was not a

result of the method of heating, but rather the performance of the formulation.

As the bottom of the particles started to melt/char, the pan was tilted to get the

material to flow the lip of the pan for possible collection in the needle. Interestingly, the

particles would move, but the portion that melted on the bottom would not flow. The

particle would slowly disintegrate as it slid along the pan, which would leave a trail, but

would stop moving once the particle completely melted. A trail can be seen in the picture

on the left hand side. The right hand side of the picture shows the portion of particles

completely melted together. After the particles on the right were melted, the pan was

slopped so that the material would flow to the right; however, the material would not pour.

As the pan was removed from the heat source, the liquefied portions became hard within a

few seconds. Because of the behavior of the material, none of it was able to be drawn

through the needle into the syringe. It was concluded that abuse by this method is not

practical.

The final step for testing abuse deterrence in this study was the water extraction

test. The results are shown in Figure 4.6.

Page 111: Copyright by Chris Eugene Brough 2015

111

Figure 4.6: Dissolution Profiles from the Water Extraction Test.

The delivery form did release slowly in ice water and room temperature water; at

the 1:15 time point there was 5% and 13% released respectively. However, the boiling

water allowed for around 40% of the drug to be released in the first 15 minutes.

Intriguingly, the multiparticulates in the boiling water appeared to have little or no physical

change as they were similar in hardness and color at the end of the test. The

multiparticulates placed in the room temperature and ice water baths had increased in size

due to water absorption, turned white in color and had taken a gelatin consistency. It seems

hydration of the formulation is key in retarding drug release and that hot water prevents

water absorption. This finding is consistent with product literature on HPMC [190, 191].

It is beyond the scope of this study to reformulate in an attempt to address the hot

water release issue. Possibly, using a lower grade of HPMC, like E3, along with a different

plasticizer or combination of plasticizers would allow for the composition to hydrate in

0.0

10.0

20.0

30.0

40.0

50.0

60.0

70.0

80.0

90.0

100.0

0:00 0:15 0:30 0:45 1:00 1:15 1:30

% D

rug

Rel

ease

Time

Boiling Water

Room Temp Water

Ice Water

Page 112: Copyright by Chris Eugene Brough 2015

112

boiling water while retaining the mechanical properties of the lead formulation. Effort on

achieving a more narrow range of particle sizes would provide a more uniform release rate

and varying the average particle size could modify the time to full drug release. Many

aspects could be investigated and optimized to improve this basic concept. However, the

study has shown that high molecular weight polymers, especially HPMC, can be

formulated for mechanical resistance to abuse methods and provide extended drug release.

Page 113: Copyright by Chris Eugene Brough 2015

113

4.5 CONCLUSION

Abuse of prescription medication is a large social challenge facing the United States

today. Abuse deterrent formulations have been introduced into the market and have shown

to discourage abusers from using those medicaments nonmedically. Some patient

populations struggle with swallowing those formulations and would benefit from a

multiparticulate abuse delivery form. This study has shown that high molecular weight

polymers, primarily HPMC, can be combined with plasticizers to create a delivery form

that is resistant to size reduction by coffee grinder, mortar and pestle and hammering. The

same formulation shows ethanol will not cause dose dumping and also resistance to

syringeability. While good water extraction resistance was established, it could be

improved upon to resist extraction in hot water. Further research could allow this concept

to be developed commercially and aid in the fight against the epidemic of prescription drug

overdose deaths.

Page 114: Copyright by Chris Eugene Brough 2015

114

Bibliography

1. Brough, C. and R.O. Williams Iii, Amorphous solid dispersions and nano-crystal

technologies for poorly water-soluble drug delivery. International Journal of

Pharmaceutics, 2013. 453(1): p. 157-166.

2. Ku, M.S. and W. Dulin, A biopharmaceutical classification-based Right-First-

Time formulation approach to reduce human pharmacokinetic variability and

project cycle time from First-In-Human to clinical Proof-Of-Concept.

Pharmaceutical Development and Technology, 2012. 17(3): p. 285-302.

3. Takagi, T., et al., A provisional biopharmaceutical classification of the top 200 oral

drug products in the United States, Great Britain, Spain, and Japan. Molecular

pharmaceutics, 2006. 3(6): p. 631-643.

4. Dahan, A., J.M. Miller, and G.L. Amidon, Prediction of solubility and permeability

class membership: provisional BCS classification of the world’s top oral drugs.

The AAPS journal, 2009. 11(4): p. 740-746.

5. Kipp, J.E., The role of solid nanoparticle technology in the parenteral delivery of

poorly water-soluble drugs. International Journal of Pharmaceutics, 2004. 284(1–

2): p. 109-122.

6. Singh, A., Z.A. Worku, and G. Van den Mooter, Oral formulation strategies to

improve solubility of poorly water-soluble drugs. Expert Opinion on Drug

Delivery, 2011. 8(10): p. 1361-1378.

7. Keserü, G.M. and G.M. Makara, The influence of lead discovery strategies on the

properties of drug candidates. Nature Reviews Drug Discovery, 2009. 8(3): p. 203-

212.

8. Lipinski, C.A., et al., Experimental and computational approaches to estimate

solubility and permeability in drug discovery and development settings. Advanced

Drug Delivery Reviews, 2012. 64, Supplement(0): p. 4-17.

9. Lipinski, C.A., Drug-like properties and the causes of poor solubility and poor

permeability. Journal of Pharmacological and Toxicological Methods, 2000. 44(1):

p. 235-249.

Page 115: Copyright by Chris Eugene Brough 2015

115

10. Serajuddin, A.T.M., Salt formation to improve drug solubility. Advanced Drug

Delivery Reviews, 2007. 59(7): p. 603-616.

11. Serajuddin, A.T.M. and M. Pudipeddi, Salt-selection strategies. Handbook of

pharmaceutical salts: properties, selection, and use. Weinheim: Wiley-VCH, 2008:

p. 135-160.

12. Stella, V.J. and Q. He, Cyclodextrins. Toxicologic Pathology, 2008. 36(1): p. 30-

42.

13. Neslihan Gursoy, R. and S. Benita, Self-emulsifying drug delivery systems (SEDDS)

for improved oral delivery of lipophilic drugs. Biomedicine & Pharmacotherapy,

2004. 58(3): p. 173-182.

14. Muchow, M., P. Maincent, and R.H. Müller, Lipid nanoparticles with a solid matrix

(SLN®, NLC®, LDC®) for oral drug delivery. Drug development and industrial

pharmacy, 2008. 34(12): p. 1394-1405.

15. Fenske, D.B., A. Chonn, and P.R. Cullis, Liposomal Nanomedicines: An Emerging

Field. Toxicologic Pathology, 2008. 36(1): p. 21-29.

16. Letchford, K. and H. Burt, A review of the formation and classification of

amphiphilic block copolymer nanoparticulate structures: micelles, nanospheres,

nanocapsules and polymersomes. European Journal of Pharmaceutics and

Biopharmaceutics, 2007. 65(3): p. 259-269.

17. Gullapalli, R.P., Soft gelatin capsules (softgels). Journal of Pharmaceutical

Sciences, 2010. 99(10): p. 4107-4148.

18. Schultheiss, N. and A. Newman, Pharmaceutical cocrystals and their

physicochemical properties. Crystal Growth and Design, 2009. 9(6): p. 2950-2967.

19. Stephenson, G.A., A. Aburub, and T.A. Woods, Physical stability of salts of weak

bases in the solid‐state. Journal of Pharmaceutical Sciences, 2011. 100(5): p. 1607-

1617.

20. Singhal, D. and W. Curatolo, Drug polymorphism and dosage form design: a

practical perspective. Advanced Drug Delivery Reviews, 2004. 56(3): p. 335-347.

Page 116: Copyright by Chris Eugene Brough 2015

116

21. Yu, L.X., et al., Biopharmaceutics classification system: the scientific basis for

biowaiver extensions. Pharmaceutical Research, 2002. 19(7): p. 921-925.

22. Wu, C.-Y. and L.Z. Benet, Predicting drug disposition via application of BCS:

transport/absorption/elimination interplay and development of a biopharmaceutics

drug disposition classification system. Pharmaceutical Research, 2005. 22(1): p.

11-23.

23. Curatolo, W., J.A. Nightingale, and S.M. Herbig, Utility of

hydroxypropylmethylcellulose acetate succinate (HPMCAS) for initiation and

maintenance of drug supersaturation in the GI milieu. Pharmaceutical Research,

2009. 26(6): p. 1419-1431.

24. Huang, L.-F. and W.-Q. Tong, Impact of solid state properties on developability

assessment of drug candidates. Advanced Drug Delivery Reviews, 2004. 56(3): p.

321-334.

25. Dressman, J.B., et al., Estimating drug solubility in the gastrointestinal tract.

Advanced Drug Delivery Reviews, 2007. 59(7): p. 591-602.

26. Zhao, Y.H., et al., Rate-limited steps of human oral absorption and QSAR studies.

Pharmaceutical Research, 2002. 19(10): p. 1446-1457.

27. Möschwitzer, J., et al., Nanocrystal formulations for improved delivery of poorly

soluble drugs. Nanomedicine in Health and Disease, 2011: p. 79.

28. Dokoumetzidis, A. and P. Macheras, A century of dissolution research: from Noyes

and Whitney to the biopharmaceutics classification system. International Journal of

Pharmaceutics, 2006. 321(1): p. 1-11.

29. Galli, C., Experimental determination of the diffusion boundary layer width of

micron and submicron particles. International Journal of Pharmaceutics, 2006.

313(1): p. 114-122.

30. Shchekin, A. and A. Rusanov, Generalization of the Gibbs–Kelvin–Köhler and

Ostwald–Freundlich equations for a liquid film on a soluble nanoparticle. The

Journal of chemical physics, 2008. 129: p. 154116.

Page 117: Copyright by Chris Eugene Brough 2015

117

31. Kesisoglou, F., S. Panmai, and Y. Wu, Nanosizing — Oral formulation

development and biopharmaceutical evaluation. Advanced Drug Delivery

Reviews, 2007. 59(7): p. 631-644.

32. Shegokar, R. and R.H. Müller, Nanocrystals: Industrially feasible multifunctional

formulation technology for poorly soluble actives. International Journal of

Pharmaceutics, 2010. 399(1–2): p. 129-139.

33. Junghanns, J.U.A.H. and R.H. Müller, Nanocrystal technology, drug delivery and

clinical applications. International journal of nanomedicine, 2008. 3(3): p. 295.

34. Choi, J.-Y., et al., Role of polymeric stabilizers for drug nanocrystal dispersions.

Current Applied Physics, 2005. 5(5): p. 472-474.

35. Van Eerdenbrugh, B., G. Van den Mooter, and P. Augustijns, Top-down production

of drug nanocrystals: Nanosuspension stabilization, miniaturization and

transformation into solid products. International Journal of Pharmaceutics, 2008.

364(1): p. 64-75.

36. Lee, J., J.Y. Choi, and C.H. Park, Characteristics of polymers enabling nano-

comminution of water-insoluble drugs. International Journal of Pharmaceutics,

2008. 355(1–2): p. 328-336.

37. Voorhees, P.W., The theory of Ostwald ripening. Journal of Statistical Physics,

1985. 38(1-2): p. 231-252.

38. Merisko-Liversidge, E., G.G. Liversidge, and E.R. Cooper, Nanosizing: a

formulation approach for poorly-water-soluble compounds. European Journal of

Pharmaceutical Sciences, 2003. 18(2): p. 113-120.

39. Hancock, B.C. and M. Parks, What is the true solubility advantage for amorphous

pharmaceuticals? Pharmaceutical Research, 2000. 17(4): p. 397-404.

40. Miller, D.A., et al., Targeted intestinal delivery of supersaturated itraconazole for

improved oral absorption. Pharmaceutical Research, 2008. 25(6): p. 1450-1459.

41. Miller, D.A., et al., Enhanced in vivo absorption of itraconazole via stabilization

of supersaturation following acidic-to-neutral pH transition. Drug development

and industrial pharmacy, 2008. 34(8): p. 890-902.

Page 118: Copyright by Chris Eugene Brough 2015

118

42. DiNunzio, J.C., et al., Amorphous compositions using concentration enhancing

polymers for improved bioavailability of itraconazole. Molecular pharmaceutics,

2008. 5(6): p. 968-980.

43. Guzmán, H.R., et al., Combined use of crystalline salt forms and precipitation

inhibitors to improve oral absorption of celecoxib from solid oral formulations.

Journal of Pharmaceutical Sciences, 2007. 96(10): p. 2686-2702.

44. Brouwers, J., M.E. Brewster, and P. Augustijns, Supersaturating drug delivery

systems: The answer to solubility-limited oral bioavailability? Journal of

Pharmaceutical Sciences, 2009. 98(8): p. 2549-2572.

45. Janssens, S. and G. Van den Mooter, Review: physical chemistry of solid

dispersions. Journal of Pharmacy and Pharmacology, 2010. 61(12): p. 1571-1586.

46. Lindfors, L., et al., Nucleation and crystal growth in supersaturated solutions of a

model drug. Journal of colloid and interface science, 2008. 325(2): p. 404-413.

47. Janssens, S., et al., Formulation and characterization of ternary solid dispersions

made up of Itraconazole and two excipients, TPGS 1000 and PVPVA 64, that were

selected based on a supersaturation screening study. European Journal of

Pharmaceutics and Biopharmaceutics, 2008. 69(1): p. 158-166.

48. Raghavan, S.L., et al., Crystallization of hydrocortisone acetate: influence of

polymers. International Journal of Pharmaceutics, 2001. 212(2): p. 213-221.

49. Raghavan, S.L., et al., Formation and stabilisation of triclosan colloidal

suspensions using supersaturated systems. International Journal of Pharmaceutics,

2003. 261(1–2): p. 153-158.

50. Vasanthavada, M., et al., Phase behavior of amorphous molecular dispersions I:

Determination of the degree and mechanism of solid solubility. Pharmaceutical

Research, 2004. 21(9): p. 1598-1606.

51. Wang, X., A. Michoel, and G. Van den Mooter, Solid state characteristics of

ternary solid dispersions composed of PVP VA64, Myrj 52 and itraconazole.

International Journal of Pharmaceutics, 2005. 303(1–2): p. 54-61.

Page 119: Copyright by Chris Eugene Brough 2015

119

52. Hancock, B.C., S.L. Shamblin, and G. Zografi, Molecular mobility of amorphous

pharmaceutical solids below their glass transition temperatures. Pharmaceutical

Research, 1995. 12(6): p. 799-806.

53. Konno, H. and L.S. Taylor, Influence of different polymers on the crystallization

tendency of molecularly dispersed amorphous felodipine. Journal of

Pharmaceutical Sciences, 2006. 95(12): p. 2692-2705.

54. Huang, J., R.J. Wigent, and J.B. Schwartz, Drug–polymer interaction and its

significance on the physical stability of nifedipine amorphous dispersion in

microparticles of an ammonio methacrylate copolymer and ethylcellulose binary

blend. Journal of Pharmaceutical Sciences, 2008. 97(1): p. 251-262.

55. Karavas, E., et al., Investigation of the release mechanism of a sparingly water-

soluble drug from solid dispersions in hydrophilic carriers based on physical state

of drug, particle size distribution and drug–polymer interactions. European Journal

of Pharmaceutics and Biopharmaceutics, 2007. 66(3): p. 334-347.

56. Chokshi, R.J., et al., Stabilization of low glass transition temperature indomethacin

formulations: Impact of polymer-type and its concentration. Journal of

Pharmaceutical Sciences, 2008. 97(6): p. 2286-2298.

57. Qian, F., J. Huang, and M.A. Hussain, Drug–polymer solubility and miscibility:

Stability consideration and practical challenges in amorphous solid dispersion

development. Journal of Pharmaceutical Sciences, 2010. 99(7): p. 2941-2947.

58. Waterman, K.C. and B.C. MacDonald, Package selection for moisture protection

for solid, oral drug products. Journal of Pharmaceutical Sciences, 2010. 99(11): p.

4437-4452.

59. Möschwitzer, J.P., Drug nanocrystals in the commercial pharmaceutical

development process. International Journal of Pharmaceutics, 2012.

60. Möschwitzer, J., Particle Size Reduction Technologies in the Pharmaceutical

Development Process. American Pharmaceutical Review, 2010 (April), 2010: p.

54-59.

Page 120: Copyright by Chris Eugene Brough 2015

120

61. Merisko-Liversidge, E.M. and G.G. Liversidge, Drug Nanoparticles: Formulating

Poorly Water-Soluble Compounds. Toxicologic Pathology, 2008. 36(1): p. 43-48.

62. Merisko-Liversidge, E. and G.G. Liversidge, Nanosizing for oral and parenteral

drug delivery: A perspective on formulating poorly-water soluble compounds using

wet media milling technology. Advanced Drug Delivery Reviews, 2011. 63(6): p.

427-440.

63. Cooper, E.R., Nanoparticles: A personal experience for formulating poorly water

soluble drugs. Journal of Controlled Release, 2010. 141(3): p. 300-302.

64. Arunkumar, N., M. Deecaraman, and C. Rani, Nanosuspension technology and its

applications in drug delivery. Asian journal of pharmaceutics, 2009. 3(3): p. 168.

65. Texter, J., Precipitation and condensation of organic particles. Journal of

dispersion science and technology, 2001. 22(6): p. 499-527.

66. de Waard, H., W.L.J. Hinrichs, and H.W. Frijlink, A novel bottom–up process to

produce drug nanocrystals: Controlled crystallization during freeze-drying.

Journal of Controlled Release, 2008. 128(2): p. 179-183.

67. Vehring, R., Pharmaceutical particle engineering via spray drying. Pharmaceutical

Research, 2008. 25(5): p. 999-1022.

68. Wu, C. and J.W. McGinity, Influence of ibuprofen as a solid-state plasticizer in

Eudragit® RS 30 D on the physicochemical properties of coated beads. AAPS

PharmSciTech, 2001. 2(4): p. 35-43.

69. Crowley, M.M., et al., Pharmaceutical applications of hot-melt extrusion: part I.

Drug development and industrial pharmacy, 2007. 33(9): p. 909-926.

70. Fini, A., et al., Ultrasound-compacted and spray-congealed

indomethacin/polyethyleneglycol systems. International Journal of Pharmaceutics,

2002. 247(1–2): p. 11-22.

71. Miller, D.A., et al., KinetiSol: A New Processing Paradigm for Amorphous Solid

Dispersion Systems. Drug Development and Delivery, 2012. 12(9).

Page 121: Copyright by Chris Eugene Brough 2015

121

72. Fakes, M.G., et al., Enhancement of oral bioavailability of an HIV-attachment

inhibitor by nanosizing and amorphous formulation approaches. International

Journal of Pharmaceutics, 2009. 370(1–2): p. 167-174.

73. Zheng, W., et al., Selection of oral bioavailability enhancing formulations during

drug discovery. Drug development and industrial pharmacy, 2012. 38(2): p. 235-

247.

74. Vogt, M., K. Kunath, and J.B. Dressman, Dissolution enhancement of fenofibrate

by micronization, cogrinding and spray-drying: Comparison with commercial

preparations. European Journal of Pharmaceutics and Biopharmaceutics, 2008.

68(2): p. 283-288.

75. Kwong, E., J. Higgins, and A.C. Templeton, Strategies for bringing drug delivery

tools into discovery. International Journal of Pharmaceutics, 2011. 412(1): p. 1-7.

76. Thombre, A.G., et al., In vitro and in vivo characterization of amorphous,

nanocrystalline, and crystalline ziprasidone formulations. International Journal of

Pharmaceutics, 2012. 428(1–2): p. 8-17.

77. Thombre, A.G., S.M. Herbig, and J.A. Alderman, Improved Ziprasidone

Formulations with Enhanced Bioavailability in the Fasted State and a Reduced

Food Effect. Pharmaceutical Research, 2011. 28(12): p. 3159-3170.

78. Sigfridsson, K., et al., A formulation comparison, using a solution and different

nanosuspensions of a poorly soluble compound. European Journal of

Pharmaceutics and Biopharmaceutics, 2007. 67(2): p. 540-547.

79. Shen, L.J., Nanomedicines in renal transplant rejection–focus on sirolimus.

International journal of nanomedicine, 2007. 2(1): p. 25-32.

80. Wu, Y., et al., The role of biopharmaceutics in the development of a clinical

nanoparticle formulation of MK-0869: a Beagle dog model predicts improved

bioavailability and diminished food effect on absorption in human. International

Journal of Pharmaceutics, 2004. 285(1–2): p. 135-146.

81. Hoy, S.M., L.J. Scott, and G.M. Keating, Intramuscular paliperidone palmitate.

CNS drugs, 2010. 24(3): p. 227-244.

Page 122: Copyright by Chris Eugene Brough 2015

122

82. Kawabata, Y., et al., Formulation design for poorly water-soluble drugs based on

biopharmaceutics classification system: Basic approaches and practical

applications. International Journal of Pharmaceutics, 2011. 420(1): p. 1-10.

83. Klein, C.E., et al., The tablet formulation of lopinavir/ritonavir provides similar

bioavailability to the soft-gelatin capsule formulation with less pharmacokinetic

variability and diminished food effect. JAIDS Journal of Acquired Immune

Deficiency Syndromes, 2007. 44(4): p. 401-410.

84. Sherman, E.M. and J.G. Steinberg, Heat-stable ritonavir tablets: a new formulation

of a pharmacokinetic enhancer for HIV. Expert Opinion on Pharmacotherapy,

2011. 12(1): p. 141-148.

85. Bauer, J., et al., Ritonavir: An Extraordinary Example of Conformational

Polymorphism. Pharmaceutical Research, 2001. 18(6): p. 859-866.

86. Chemburkar, S.R., et al., Dealing with the Impact of Ritonavir Polymorphs on the

Late Stages of Bulk Drug Process Development. Organic Process Research &

Development, 2000. 4(5): p. 413-417.

87. Verreck, G., et al., Characterization of solid dispersions of itraconazole and

hydroxypropylmethylcellulose prepared by melt extrusion—part I. International

Journal of Pharmaceutics, 2003. 251(1–2): p. 165-174.

88. Lee, S., et al., Preparation and characterization of solid dispersions of itraconazole

by using aerosol solvent extraction system for improvement in drug solubility and

bioavailability. Archives of Pharmacal Research, 2005. 28(7): p. 866-874.

89. Six, K., et al., Characterization of solid dispersions of itraconazole and

hydroxypropylmethylcellulose prepared by melt extrusion, part II. Pharmaceutical

Research, 2003. 20(7): p. 1047-1054.

90. Shah, N., et al., Development of Novel Microprecipitated Bulk Powder (MBP)

Technology for Manufacturing Stable Amorphous Formulations of Poorly Soluble

Drugs. International Journal of Pharmaceutics, 2012.

91. Harmon, A., A Roller Coaster Chase for a Cure, in The New York Times2010.

Page 123: Copyright by Chris Eugene Brough 2015

123

92. Harmon, A., After Long Fight, Drug Gives Sudden Reprieve, in The New York

Times2010.

93. Van Drooge, D., et al., Characterization of the molecular distribution of drugs in

glassy solid dispersions at the nano-meter scale, using differential scanning

calorimetry and gravimetric water vapour sorption techniques. International

Journal of Pharmaceutics, 2006. 310(1): p. 220-229.

94. Verreck, G., et al., The effect of pressurized carbon dioxide as a temporary

plasticizer and foaming agent on the hot stage extrusion process and extrudate

properties of solid dispersions of itraconazole with PVP-VA 64. European Journal

of Pharmaceutical Sciences, 2005. 26(3): p. 349-358.

95. Shibata, Y., et al., Effect of characteristics of compounds on maintenance of an

amorphous state in solid dispersion with crospovidone. Journal of Pharmaceutical

Sciences, 2007. 96(6): p. 1537-1547.

96. Urbanetz, N.A., Stabilization of solid dispersions of nimodipine and polyethylene

glycol 2000. European Journal of Pharmaceutical Sciences, 2006. 28(1): p. 67-76.

97. Janssens, S., et al., The use of a new hydrophilic polymer, Kollicoat IR<

sup>®</sup>, in the formulation of solid dispersions of Itraconazole. European

Journal of Pharmaceutical Sciences, 2007. 30(3): p. 288-294.

98. Albers, J., et al., Mechanism of drug release from polymethacrylate-based

extrudates and milled strands prepared by hot-melt extrusion. European Journal of

Pharmaceutics and Biopharmaceutics, 2009. 71(2): p. 387-394.

99. Chokshi, R.J., et al., Characterization of physico-mechanical properties of

indomethacin and polymers to assess their suitability for hot-melt extrusion

processs as a means to manufacture solid dispersion/solution. Journal of

Pharmaceutical Sciences, 2005. 94(11): p. 2463-2474.

100. Tanno, F., et al., Evaluation of hypromellose acetate succinate (HPMCAS) as a

carrier in solid dispersions. Drug development and industrial pharmacy, 2004.

30(1): p. 9-17.

Page 124: Copyright by Chris Eugene Brough 2015

124

101. Friesen, D.T., et al., Hydroxypropyl methylcellulose acetate succinate-based spray-

dried dispersions: an overview. Molecular pharmaceutics, 2008. 5(6): p. 1003-

1019.

102. Ghosh, I., et al., Comparison of HPMC based polymers performance as carriers

for manufacture of solid dispersions using the melt extruder. International Journal

of Pharmaceutics, 2011. 419(1): p. 12-19.

103. Chowdary, K. and K. Suresh Babu, Dissolution, bioavailability and ulcerogenic

studies on solid dispersions of indomethacin in water soluble cellulose polymers.

Drug development and industrial pharmacy, 1994. 20(5): p. 799-813.

104. Onoue, S., et al., Improved dissolution and pharmacokinetic behavior of

cyclosporine A using high-energy amorphous solid dispersion approach.

International Journal of Pharmaceutics, 2010. 399(1): p. 94-101.

105. Newa, M., et al., Preparation, characterization and in vivo evaluation of ibuprofen

binary solid dispersions with poloxamer 188. International Journal of

Pharmaceutics, 2007. 343(1): p. 228-237.

106. DiNunzio, J.C., et al., Applications of KinetiSol< sup>®</sup> Dispersing for the

production of plasticizer free amorphous solid dispersions. European Journal of

Pharmaceutical Sciences, 2010. 40(3): p. 179-187.

107. Djuris, J., et al., Preparation of carbamazepine–Soluplus® solid dispersions by hot-

melt extrusion, and prediction of drug–polymer miscibility by thermodynamic

model fitting. European Journal of Pharmaceutics and Biopharmaceutics, 2013.

84(1): p. 228-237.

108. Rowe, R.C., et al., Handbook of pharmaceutical excipients. Vol. 6. 2006:

Pharmaceutical press London.

109. Alcohol, M.-P., Technical Brochure, 1999, Clariant GmgH: Sulzbach, Germany.

110. Galindo-Rodriguez, S., et al., Physicochemical parameters associated with

nanoparticle formation in the salting-out, emulsification-diffusion, and

nanoprecipitation methods. Pharmaceutical Research, 2004. 21(8): p. 1428-1439.

Page 125: Copyright by Chris Eugene Brough 2015

125

111. Abdel-Mottaleb, M.M., et al., Physically cross-linked polyvinyl alcohol for the

topical delivery of fluconazole. Drug development and industrial pharmacy, 2009.

35(3): p. 311-320.

112. Bourges, J.L., et al., Intraocular implants for extended drug delivery: Therapeutic

applications. Advanced Drug Delivery Reviews, 2006. 58(11): p. 1182-1202.

113. Davies, N.M., et al., Evaluation of mucoadhesive polymers in ocular drug delivery.

I. Viscous solutions. Pharmaceutical Research, 1991. 8(8): p. 1039-1043.

114. McDonald, C., et al., A randomised, crossover, multicentre study to compare the

performance of 0.1&percnt;(w/v) sodium hyaluronate with 1.4&percnt;(w/v)

polyvinyl alcohol in the alleviation of symptoms associated with dry eye syndrome.

Eye, 2002. 16(5): p. 601-607.

115. Winterton, L.C., et al., The elution of poly (vinyl alcohol) from a contact lens: The

realization of a time release moisturizing agent/artificial tear. Journal of

Biomedical Materials Research Part B: Applied Biomaterials, 2007. 80B(2): p.

424-432.

116. Thanoo, B., M. Sunny, and A. Jayakrishnan, Controlled Release of Oral Drugs

from Cross‐linked Polyvinyl Alcohol Microspheres. Journal of Pharmacy and

Pharmacology, 1993. 45(1): p. 16-20.

117. Sahoo, S.K., et al., Residual polyvinyl alcohol associated with poly (D, L-lactide-

co-glycolide) nanoparticles affects their physical properties and cellular uptake.

Journal of Controlled Release, 2002. 82(1): p. 105-114.

118. Mu, L. and S. Feng, A novel controlled release formulation for the anticancer drug

paclitaxel (Taxol< sup>®</sup>): PLGA nanoparticles containing vitamin E

TPGS. Journal of Controlled Release, 2003. 86(1): p. 33-48.

119. Riis, T., et al., pH-independent drug release of an extremely poorly soluble weakly

acidic drug from multiparticulate extended release formulations. European Journal

of Pharmaceutics and Biopharmaceutics, 2007. 65(1): p. 78-84.

120. Davaran, S., et al., Development of a novel prolonged-release nicotine transdermal

patch. Pharmacological research, 2005. 51(3): p. 233-237.

Page 126: Copyright by Chris Eugene Brough 2015

126

121. DeMerlis, C. and D. Schoneker, Review of the oral toxicity of polyvinyl alcohol

(PVA). Food and Chemical Toxicology, 2003. 41(3): p. 319-326.

122. De Jaeghere, W., et al., HOT-MELT EXTRUSION OF POLYVINYL ALCOHOL

FOR ORAL IMMEDIATE RELEASE APPLICATIONS. International Journal of

Pharmaceutics, 2015.

123. Keen, J.M., J.W. McGinity, and R.O. Williams III, Enhancing bioavailability

through thermal processing. International Journal of Pharmaceutics, 2013. 450(1):

p. 185-196.

124. Gray, V., et al., The science of USP 1 and 2 dissolution: present challenges and

future relevance. Pharmaceutical Research, 2009. 26(6): p. 1289-1302.

125. Azarmi, S., W. Roa, and R. Löbenberg, Current perspectives in dissolution testing

of conventional and novel dosage forms. International Journal of Pharmaceutics,

2007. 328(1): p. 12-21.

126. Galia, E., et al., Evaluation of various dissolution media for predicting in vivo

performance of class I and II drugs. Pharmaceutical Research, 1998. 15(5): p. 698-

705.

127. Hughey, J.R., et al., Dissolution enhancement of a drug exhibiting thermal and

acidic decomposition characteristics by fusion processing: a comparative study of

hot melt extrusion and kinetisol® dispersing. AAPS PharmSciTech, 2010. 11(2):

p. 760-774.

128. Sarode, A.L., et al., Hot melt extrusion (HME) for amorphous solid dispersions:

predictive tools for processing and impact of drug–polymer interactions on

supersaturation. European Journal of Pharmaceutical Sciences, 2013. 48(3): p.

371-384.

129. Trey, S.M., et al., Delivery of itraconazole from extruded HPC films. Drug

development and industrial pharmacy, 2007. 33(7): p. 727-735.

130. Nunes, C., A. Mahendrasingam, and R. Suryanarayanan, Quantification of

crystallinity in substantially amorphous materials by synchrotron X-ray powder

diffractometry. Pharmaceutical Research, 2005. 22(11): p. 1942-1953.

Page 127: Copyright by Chris Eugene Brough 2015

127

131. Wang, B., et al., Viscometric, light scattering, and size-exclusion chromatography

studies on the structural changes of aqueous poly(vinyl alcohol) induced by γ-ray

irradiation. Journal of Polymer Science Part B: Polymer Physics, 2000. 38(1): p.

214-221.

132. Hughey, J.R., et al., Preparation and characterization of fusion processed solid

dispersions containing a viscous thermally labile polymeric carrier. International

Journal of Pharmaceutics, 2012. 438(1): p. 11-19.

133. Hutchinson, J.M. and P. Kumar, Enthalpy relaxation in polyvinyl acetate.

Thermochimica acta, 2002. 391(1): p. 197-217.

134. Lu, J., T. Wang, and L.T. Drzal, Preparation and properties of microfibrillated

cellulose polyvinyl alcohol composite materials. Composites Part A: Applied

Science and Manufacturing, 2008. 39(5): p. 738-746.

135. Stoclet, G., R. Seguela, and J.-M. Lefebvre, Morphology, thermal behavior and

mechanical properties of binary blends of compatible biosourced polymers:

Polylactide/polyamide11. Polymer, 2011. 52(6): p. 1417-1425.

136. Richardson, M., P. Flory, and J. Jackson, Crystallization and melting of copolymers

of polymethylene. Polymer, 1963. 4: p. 221-236.

137. Six, K., et al., Investigation of thermal properties of glassy itraconazole:

identification of a monotropic mesophase. Thermochimica acta, 2001. 376(2): p.

175-181.

138. Chokshi, R.J., et al., Characterization of physico‐mechanical properties of

indomethacin and polymers to assess their suitability for hot‐melt extrusion

processs as a means to manufacture solid dispersion/solution. Journal of

Pharmaceutical Sciences, 2005. 94(11): p. 2463-2474.

139. Rambali, B., et al., Itraconazole formulation studies of the melt-extrusion process

with mixture design. Drug development and industrial pharmacy, 2003. 29(6): p.

641-652.

Page 128: Copyright by Chris Eugene Brough 2015

128

140. Vasconcelos, T., B. Sarmento, and P. Costa, Solid dispersions as strategy to

improve oral bioavailability of poor water soluble drugs. Drug Discovery Today,

2007. 12(23–24): p. 1068-1075.

141. Vo, C.L.-N., C. Park, and B.-J. Lee, Current trends and future perspectives of solid

dispersions containing poorly water-soluble drugs. European Journal of

Pharmaceutics and Biopharmaceutics, 2013. 85(3): p. 799-813.

142. Van den Mooter, G., The use of amorphous solid dispersions: A formulation

strategy to overcome poor solubility and dissolution rate. Drug Discovery Today:

Technologies, 2012. 9(2): p. e79-e85.

143. Six, K., et al., Clinical study of solid dispersions of itraconazole prepared by hot-

stage extrusion. European Journal of Pharmaceutical Sciences, 2005. 24(2): p. 179-

186.

144. Mellaerts, R., et al., Increasing the oral bioavailability of the poorly water soluble

drug itraconazole with ordered mesoporous silica. European Journal of

Pharmaceutics and Biopharmaceutics, 2008. 69(1): p. 223-230.

145. Miller, D.A., et al., Targeted intestinal delivery of supersaturated itraconazole for

improved oral absorption. Pharmaceutical Research, 2008. 25(6): p. 1450-1459.

146. DiNunzio, J.C., et al., Production of advanced solid dispersions for enhanced

bioavailability of itraconazole using KinetiSol® Dispersing. Drug development

and industrial pharmacy, 2010. 36(9): p. 1064-1078.

147. Zhang, K., et al., Increased dissolution and oral absorption of

itraconazole/Soluplus extrudate compared with itraconazole nanosuspension.

European Journal of Pharmaceutics and Biopharmaceutics, 2013. 85(3): p. 1285-

1292.

148. Law, D., et al., Physicochemical considerations in the preparation of amorphous

ritonavir–poly(ethylene glycol) 8000 solid dispersions. Journal of Pharmaceutical

Sciences, 2001. 90(8): p. 1015-1025.

Page 129: Copyright by Chris Eugene Brough 2015

129

149. Bley, H., B. Fussnegger, and R. Bodmeier, Characterization and stability of solid

dispersions based on PEG/polymer blends. International Journal of Pharmaceutics,

2010. 390(2): p. 165-173.

150. Newa, M., et al., Preparation, characterization and in vivo evaluation of ibuprofen

binary solid dispersions with poloxamer 188. International Journal of

Pharmaceutics, 2007. 343(1): p. 228-237.

151. Craig, D.Q., The mechanisms of drug release from solid dispersions in water-

soluble polymers. International Journal of Pharmaceutics, 2002. 231(2): p. 131-144.

152. Kazarian, S. and G. Martirosyan, Spectroscopy of polymer/drug formulations

processed with supercritical fluids: in situ ATR–IR and Raman study of

impregnation of ibuprofen into PVP. International Journal of Pharmaceutics, 2002.

232(1): p. 81-90.

153. van der Weerd, J. and S.G. Kazarian, Combined approach of FTIR imaging and

conventional dissolution tests applied to drug release. Journal of Controlled

Release, 2004. 98(2): p. 295-305.

154. Nesseem, D.I., Formulation and evaluation of itraconazole via liquid crystal for

topical delivery system. Journal of pharmaceutical and biomedical analysis, 2001.

26(3): p. 387-399.

155. Garg, A., R. Sachdeva, and G. Kapoor, Comparison of crystalline and amorphous

carriers to improve the dissolution profile of water insoluble drug itraconazole. Int.

J. Pharm. Bio. Sci. 2013; 4 (1): 934, 2013. 948.

156. Calderini, A. and F. Pessine, Synthesis and characterization of inclusion complex

of the vasodilator drug minoxidil with β-cyclodextrin. Journal of Inclusion

Phenomena and Macrocyclic Chemistry, 2008. 60(3-4): p. 369-377.

157. Crescenzi, V., et al., Synthesis and partial characterization of hydrogels obtained

via glutaraldehyde crosslinking of acetylated chitosan and of hyaluronan

derivatives. Biomacromolecules, 2003. 4(4): p. 1045-1054.

Page 130: Copyright by Chris Eugene Brough 2015

130

158. Trefi, S., et al., The usefulness of 2D DOSY and 3D DOSY‐COSY 1H NMR for

mixture analysis: application to genuine and fake formulations of sildenafil

(Viagra). Magnetic Resonance in Chemistry, 2009. 47(S1): p. S163-S173.

159. Laurienzo, P., et al., Novel alginate–acrylic polymers as a platform for drug

delivery. Journal of Biomedical Materials Research Part A, 2006. 78(3): p. 523-

531.

160. Ambike, A.A., K. Mahadik, and A. Paradkar, Stability study of amorphous

valdecoxib. International Journal of Pharmaceutics, 2004. 282(1): p. 151-162.

161. Office, U.S.G.A., Prescription Pain Reliever Abuse. Report to Congressional

Requesters, 2011.

162. Prevention, C.f.D.C.a., Unintentional Drug Poisoning in the United States. CDC

Brief, 2010.

163. Abuse, S. and M. Administration, Results from the 2010 national survey on drug

use and health: Summary of national findings, NSDUH Series H-41, HHS

Publication No.(SMA) 11-4658, in Substance Abuse and Mental Health Services

Administration2011, Rockville, MD.

164. Food, U. and D. Administration, Guidance for industry abuse-deterrent opioids-

evaluation and labeling. Center for Drug Evaluation and Research, 2013.

165. Ma, L., L. Deng, and J. Chen, Applications of poly (ethylene oxide) in controlled

release tablet systems: a review. Drug development and industrial pharmacy, 2014.

40(7): p. 845-851.

166. Alexander, L., et al., Development and impact of prescription opioid abuse

deterrent formulation technologies. Drug and alcohol dependence, 2014. 138: p. 1-

6.

167. Cicero, T.J., M.S. Ellis, and H.L. Surratt, Effect of abuse-deterrent formulation of

OxyContin. New England Journal of Medicine, 2012. 367(2): p. 187-189.

168. Butler, S.F., et al., Abuse rates and routes of administration of reformulated

extended-release oxycodone: initial findings from a sentinel surveillance sample of

Page 131: Copyright by Chris Eugene Brough 2015

131

individuals assessed for substance abuse treatment. The Journal of Pain, 2013.

14(4): p. 351-358.

169. Cone, E.J., J. Giordano, and B. Weingarten, An iterative model for in vitro

laboratory assessment of tamper deterrent formulations. Drug and alcohol

dependence, 2013. 131(1): p. 100-105.

170. Harris, S.C., et al., Abuse potential, pharmacokinetics, pharmacodynamics, and

safety of intranasally administered crushed oxycodone HCl abuse‐deterrent

controlled‐release tablets in recreational opioid users. The Journal of Clinical

Pharmacology, 2014. 54(4): p. 468-477.

171. Raffa, R.B. and J.V. Pergolizzi Jr, Opioid formulations designed to resist/deter

abuse. Drugs, 2010. 70(13): p. 1657-1675.

172. Butler, S.F., et al., Estimating Attractiveness for Abuse of a Not‐Yet‐Marketed

“Abuse‐Deterrent” Prescription Opioid Formulation. Pain Medicine, 2010. 11(1):

p. 81-91.

173. Merrigan, T., et al. Abuse-resistant, controlled-release oxycodone treats chronic

pain. in American Academy of Pain Medicine 24th Annual Meeting. 2008.

174. Barry, D.T., et al., Pain and associated substance use among opioid dependent

individuals seeking office‐based treatment with buprenorphine–naloxone: A needs

assessment study. The American Journal on Addictions, 2013. 22(3): p. 212-217.

175. Duggan, S.T. and L.J. Scott, Morphine/naltrexone. CNS drugs, 2010. 24(6): p. 527-

538.

176. Engel, K., et al., Anti‐inflammatory effect of pimecrolimus in the sodium lauryl

sulphate test. Journal of the European Academy of Dermatology and Venereology,

2008. 22(4): p. 447-450.

177. Andersen, O., et al., [Problems when swallowing tablets. A questionnaire study

from general practice]. Tidsskrift for den Norske laegeforening: tidsskrift for

praktisk medicin, ny raekke, 1995. 115(8): p. 947-949.

178. Aldworth, J., Results from the 2007 national survey on drug use and health:

National findings2009: DIANE Publishing.

Page 132: Copyright by Chris Eugene Brough 2015

132

179. Von Korff, M., et al., An epidemiologic comparison of pain complaints. Pain, 1988.

32(2): p. 173-183.

180. Bouhassira, D., et al., Prevalence of chronic pain with neuropathic characteristics

in the general population. Pain, 2008. 136(3): p. 380-387.

181. Kopecky, A., A.B. Fleming, and P.K. Noonan, Safety and pharmacokinetics of

oxycodone DETERx administered intranasally in recreational opioid users. Drug

& Alcohol Dependence, 2015. 146: p. e163.

182. Gudin, J., et al., Comparing the Effect of Tampering on the Oral Pharmacokinetic

Profiles of Two Extended‐Release Oxycodone Formulations with Abuse‐Deterrent

Properties. Pain Medicine, 2015.

183. Fleming, A.B., et al., Evaluation of an Extended‐Release, Abuse‐Deterrent,

Microsphere‐in‐Capsule Analgesic for the Management of Patients with Chronic

Pain With Dysphagia (CPD). Pain Practice, 2015.

184. Coppens, K.A., et al., Hypromellose, ethylcellulose, and polyethylene oxide use in

hot melt extrusion. 2006.

185. Hale, M.E., et al., Efficacy and safety of controlled-release versus immediate-

release oxycodone: randomized, double-blind evaluation in patients with chronic

back pain. The Clinical Journal of Pain, 1999. 15(3): p. 179-183.

186. Saltzman, W.M. and W.L. Olbricht, Building drug delivery into tissue engineering

design. Nature Reviews Drug Discovery, 2002. 1(3): p. 177-186.

187. Utracki, L.A., Compatibilization of polymer blends. 2002.

188. Crowley, M.M., et al., Physicochemical properties and mechanism of drug release

from ethyl cellulose matrix tablets prepared by direct compression and hot-melt

extrusion. International Journal of Pharmaceutics, 2004. 269(2): p. 509-522.

189. Shah, V.P., et al., In vitro dissolution profile comparison—statistics and analysis

of the similarity factor, f2. Pharmaceutical Research, 1998. 15(6): p. 889-896.

190. Rajabi-Siahboomi, A., et al., Structure and behaviour in hydrophilic matrix

sustained release dosage forms: 2. NMR-imaging studies of dimensional changes

Page 133: Copyright by Chris Eugene Brough 2015

133

in the gel layer and core of HPMC tablets undergoing hydration. Journal of

Controlled Release, 1994. 31(2): p. 121-128.

191. Sarkar, N., Kinetics of thermal gelation of methylcellulose and

hydroxypropylmethylcellulose in aqueous solutions. Carbohydrate Polymers, 1995.

26(3): p. 195-203.

Page 134: Copyright by Chris Eugene Brough 2015

134

Vita

Chris Eugene Brough graduated from high school at Hong Kong International

School in 1991. After which he attended Brigham Young University in Provo, Utah. He

received B.S. and M.S. degrees in mechanical engineering and also completed a Master’s

degree in Business Administration from the Marriott School of Management. Upon

completing his education, he formed a company called ReSyk to exploit the technology he

developed in his master’s thesis.

ReSyk sold machines, called compounders, to companies that were looking to

recycle plastic directly into commodity products. ReSyk had 7 licensees of the technology

within the United States and Mexico when it sold to IntegriCo Composites in 2005.

IntegriCo moved to Temple, Texas, in 2006 to set up a large manufacturing facility to

produce plastic railroad ties from recycled post-industrial and post-consumer plastic. At

this time, Chris came into contact with Dave Miller, who was a Ph.D. student at the

University of Texas. Dave was trying to produce viscous, pharmaceutical amorphous

dispersions via hot melt extrusion with limited success. Chris and Dave tested a few of

these compositions in the large industrial compounders and had immediate success.

With this result, IntegriCo decided to form a spin-off company, called DisperSol

Technologies, to develop the technology for the pharmaceutical fields. Chris built a small,

lab scale compounder and placed it at the University in 2008 and worked with Ph.D.

candidates on the technology. At DisperSol’s persistent request, Chris became a student

in 2011 to explore new areas with the technology.

Email address: [email protected]

This dissertation was typed by the author.